You are on page 1of 50

3.2 5.

Review

Green Hydrogen Production


Technologies from Ammonia
Cracking

Hossein Ali Yousefi Rizi and Donghoon Shin

Special Issue
Heat Conversion and Emission Characteristics in Fuel Combustion Processes
Edited by
Prof. Dr. Donghoon Shin and Dr. Hossein Ali Yousefi Rizi

https://doi.org/10.3390/en15218246
energies
Review
Green Hydrogen Production Technologies from
Ammonia Cracking
Hossein Ali Yousefi Rizi 1 and Donghoon Shin 2, *

1 School of Mechanical and Automotive Engineering, Kookmin University, Seoul 02707, Korea
2 Department of Mechanical Engineering, Kookmin University, Seoul 02707, Korea
* Correspondence: d.shin@kookmin.ac.kr

Abstract: The rising technology of green hydrogen supply systems is expected to be on the horizon.
Hydrogen is a clean and renewable energy source with the highest energy content by weight among
the fuels and contains about six times more energy than ammonia. Meanwhile, ammonia is the
most popular substance as a green hydrogen carrier because it does not carry carbon, and the
total hydrogen content of ammonia is higher than other fuels and is thus suitable to convert to
hydrogen. There are several pathways for hydrogen production. The considered aspects herein
include hydrogen production technologies, pathways based on the raw material and energy sources,
and different scales. Hydrogen can be produced from ammonia through several technologies, such
as electrochemical, photocatalytic and thermochemical processes, that can be used at production
plants and fueling stations, taking into consideration the conversion efficiency, reactors, catalysts and
their related economics. The commercial process is conducted by using expensive Ru catalysts in the
ammonia converting process but is considered to be replaced by other materials such as Ni, Co, La,
and other perovskite catalysts, which have high commercial potential with equivalent activity for
extracting hydrogen from ammonia. For successful engraftment of ammonia to hydrogen technology
into industry, integration with green technologies and economic methods, as well as safety aspects,
should be carried out.

Citation: Yousefi Rizi, H.A.; Shin, D. Keywords: hydrogen; ammonia; cracking; catalyst
Green Hydrogen Production
Technologies from Ammonia
Cracking. Energies 2022, 15, 8246.
https://doi.org/10.3390/en15218246 1. Introduction
Academic Editor: Attilio Converti Hydrogen is a clean fuel and energy carrier playing an important role in a sustainable
energy future. When consumed in a fuel cell, the energy content is converted with high
Received: 30 September 2022 efficiency on demand and produces water. Hydrogen can easily integrate into the existing
Accepted: 25 October 2022
natural gas network with minimal modification but lacks a wide infrastructure, making its
Published: 4 November 2022
storage and transport difficult and expensive [1]. Hydrogen technology for transition to a
Publisher’s Note: MDPI stays neutral hydrogen-based economy requires supplying clean and renewable energy and capture of
with regard to jurisdictional claims in CO2 from current fossil hydrogen production. The ammonia and hydrogen industry has
published maps and institutional affil- the potential to make significant contributions to global economies with the possibility of
iations. using ammonia and hydrogen as fuels [2].
This review study aims to confirm the direction of commercialization by increasing
ammonia to hydrogen production systems in the future by examining the trends of the latest
technologies of ammonia cracking, hydrogen separation and purification as a clean fuel and
Copyright: © 2022 by the authors.
source of energy without emitting greenhouse gases. The conversion technology of ammonia
Licensee MDPI, Basel, Switzerland.
to hydrogen considers the feedstock, decomposition methods, emissions, and energy efficiency,
This article is an open access article
as well as reactors and catalyst materials, with a focus on ammonia cracking.
distributed under the terms and
This paper provides a brief and comprehensive view of hydrogen production from
conditions of the Creative Commons
Attribution (CC BY) license (https://
ammonia and is organized as follows: In the first section, significant literature related to
creativecommons.org/licenses/by/
the hydrogen industry outlook is reviewed. In Section 2, hydrogen properties, production,
4.0/).

Energies 2022, 15, 8246. https://doi.org/10.3390/en15218246 https://www.mdpi.com/journal/energies


Energies 2022, 15, 8246 2 of 49

sources, and pathways are defined. Then, Section 3 is a brief look at ammonia produc-
tion technologies. Section 4 discusses hydrogen production technologies, thermal and
thermochemical cracking, and reactor technology. Section 5 reviews catalyst characteriza-
tion, activity and performance. Section 6 is a brief look at other ammonia decomposition
methods and also the separation and purification systems processes.

1.1. International Hydrogen Market Size


There are diverse applications of hydrogen being recognized worldwide. Therefore,
its demand and use are increasing across a wide range, with the potential for significant
contribution to clean energy transitions. The demand for hydrogen has increased linearly
in the past. The various applications of hydrogen as a feedstock and fuel. In the chemical
industry hydrogen is used for oil refining and the production of ammonia, methanol,
fertilizers, in food and pharmaceutical production, metal manufacturing, and low-carbon
emission steel [2,3]. Also used as fuel to produce electricity in fuel cells, and is being
adopted as a sustainable fuel for power generation in the building and transport sectors in
the future [2,3].
The annual global demand for hydrogen (in megatons) as a fuel, based on sectors
such as oil refining, industry, transport, power, ammonia fuel, synfuels, buildings, and
electric grid injection, is represented in Table 1, and Figure 1 shows the expected hydrogen
demands (%) in some sectors compared to other fuels. In the EU, expected demands will
exceed more than 100 million tons by 2050 (Figure 2). Hydrogen being produced based on
carbon capture, utilization and storage (CCUS) technology, is expected to reach net-zero
emissions by 2050 [4]. Hydrogen demand growth in new sectors, as expected in the EU,
will exceed more than 100 million tons by 2050 (Figure 2). Thus, industries are interested in
research on large-scale hydrogen projects in several regions [5].

(a) (b)

Figure 1. Expected annual demand for hydrogen based on (a) hydrogen production technologies and
(b) sectors (Mt). Hydrogen produced by electrolysis, and also as a by-product of catalytic naphtha
reforming (CNR) based on technology of carbon capture, utilization and storage (CCUS). Reprinted
with permission [4].
Energies 2022, 15, 8246 3 of 49

Table 1. Expected global hydrogen demand (in megatons) of the different energy sectors in the Net
Zero Scenario, 2020–2030. Reprinted with permission [4,5].

Grid
Refining Industry Transport Ammonia Synfuels Buildings
Power (Mt) Injection
(Mt) (Mt) (Mt) fuel (Mt) (Mt) (Mt)
(Mt)
2020 37.18 51.3 0.02 0 0 0 0.01 0
2025 33.82 63.22 2.12 0 7.53 1.1 2.25 23.85
2030 25.78 75 8.55 18.5 18.11 7.28 5.64 51.7

The potential global market for hydrogen technology is expected to reach USD
2.5 trillion [6]. Annual power generation from fuel cells has increased 15-fold since 2015,
now exceeding 1 gigawatt (1 GW) [7]. For example, the US hydrogen economy is expected
to generate approximately USD 750 billion in annual sales and 3.4 million cumulative
jobs by 2050 [7]. However, there is limited information available about the economic
competitiveness of hydrogen system configurations [8,9].

(a) (b)

Figure 2. Hydrogen prospects of the different industrial sectors in the EU, (a) current usage of
hydrogen, (b) expected hydrogen demand. * excluding potential of hydrogen in enabling renewable
energy and in heating for build gas. Reprinted with permission [10].

1.2. Hydrogen Industry Outlook


To move toward a low-carbon economy, hydrogen as an energy vector can play a
significant role in clean energy transitions, such as the release of stored chemical energy
that produces power while eliminating greenhouse gas (GHG) emissions. The utilization
of hydrogen in the power, building and transportation sectors for eco-friendly energy
production can decrease the environmental impact of emissions. Furthermore, it can be used
as a marine energy source, which is cheaper than ammonia for marine transportation [10,11].
Hydrogen technology has the technical feasibility, sufficient capacity, and potential
to participate in the electric power sector, as it can be stored and used for producing
several gigawatts per hour, enabling electrical grid stability to meet the varying demands
of different energy sectors. It can increase the utilization of generators, including nuclear,
coal, natural gas and renewables. In conjunction with fuel cells or combustion-based
technologies, it enables zero or near-zero emissions in transportation, stationary or remote
power, and portable power applications. There is also a growing interest in hydrogen fuel
cells in the aviation industry instead of heavy batteries [4,8,12–14].
Energies 2022, 15, 8246 4 of 49

The energy required to provide heating, cooling, drying, power, and fresh water to
buildings is mainly provided by fossil fuels and traditional power plants and is higher
than the transportation sector, with significant CO2 emissions associated. Hydrogen can
effectively provide the energy demand of buildings in a reliable, safe, and economical way
to reduce the global warming trend [1]. Moreover, hydrogen-powered vehicles that were
developed as a policy action to move toward a low-carbon economy are currently using
gray hydrogen [15,16]. There are some challenges in the transition from fossil-based to
renewable energy sources; for example, among carbon-free energy carriers, water, wind,
and solar are limited by geographical conditions, and produced hydrogen, needs to be
compressed, and liquified, then requires special delivery and storage systems [12,17].
The development of hydrogen as a major carbon-free energy source is a common
goal pursued by many countries to deal with the global warming trend. Hydrogen is
the most feasible energy carrier for future energy systems, as shown in Figure 3 [18,19].
The promotion of clean technologies for the production of clean hydrogen and hydrogen-
based fuels reflected in the Net Zero Emissions Scenario in 2021–2050 will reduce 60 Gt
of CO2 emissions, which is about 6% of the total cumulative emissions reduction [20,21].
Promoting eco-friendly hydrogen production and usage is necessary to reduce greenhouse
gas (GHG) emissions [3], which needs a demand generation strategy, greater efforts for
new applications, quality improvement of hydrogen production infrastructure, transitional
energy process capacity, reduced associated emissions and improved cost-effectiveness of
hydrogen production [22].

Figure 3. Transition from fossil-based to hydrogen-based energy Reprinted with permission [19].

2. Hydrogen Source, Production Pathways, Technology, and Energy Efficiency


Hydrogen production can be considered from the view of feedstocks, energy sources,
pathways, and technology.

2.1. Properties of Hydrogen


Hydrogen is a renewable, non-toxic, and carbon-free fuel that is expected to contribute
significantly to air quality improvement. Its properties are compared with ammonia and
other hydrocarbon fuels in Table 2 and Figure 4 [23,24].


Energies 2022, 15, 8246 5 of 49

Table 2. Comparing hydrogen properties with other fuels. Reprinted with permission [23,24].

Fuel Liquid H2 Gaseous H2 Natural Gas Ammonia Propane Gasoline Methanol


Formula H2 H2 CH4 NH3 C3 H8 C8 H18 CH3 OH
Storage method Cryogenic liquid Compressed gas Compressed gas Liquid Liquid Liquid Liquid
Approximate AKI * RON > 130 MON RON > 130 MON
107 110 103 87–93 113
Octane rating very low very low
Storage temp (◦ C) −253 25 25 25 25 25 25
Storage pressure (kPa) 102 24,821 24,821 1030 1020 101.3 101.3
Fuel density (kg/m3 ) 71.1 17.5 187.2 602.8 492.6 698.3 786.3
Energy storage
LHV (MJ/kg) 120.1 120.1 38.1 18.8 45.8 42.5 19.7
LHV (MJ/L) 8.5 2.1 7.1 11.3 22.6 29.7 15.5
Fuel requirement to match the energy of 10 gallons of gasoline
Fuel volume (L) 131.5 534.4 157.5 99.2 49.8 37.9 72.5
Fuel weight (kg) 9.4 9.4 29.5 59.8 24.5 26.4 57.0
AKI *: anti-knock index is the minimum octane rating for unleaded motor fuel and the mean of RON (research
octane number) and MON (motor octane number).

Hydrogen can be compressed at about 350 to 700 bar, liquefied at a low boiling point
of −253 ◦ C, or stored on the surface or within a solid. However, the current storage of
hydrogen is difficult, energy-intensive and expensive compared to the storage of ammonia
due to its high volatility, very low density and need for high pressure [18].
Hydrogen combustion with oxygen only emits water, while ammonia combustion
produces NOX , owing to the fuel NOX process. The energy required to initiate hydro-
gen combustion is much lower than that required for other common fuels, which makes
it easy to ignite. It has a very wide flammability range (between 4% and 75% in air)
compared to other fuels and a high combustion heat (3.37 MJ/kg). However, at low con-

centrations of hydrogen in air, the energy required to initiate combustion is similar to
that of other fuels [25,26]. Low volumetric energy density, low flash point, and a large
diffusion coefficient are the defects of hydrogen with regard to fuel utilization. The auto-
ignition temperature of hydrogen (585 ◦ C) is similar to natural gas, which is higher than
the gasoline vapor. Its lower heating value is 9.9 (MJ/m3 ), which is much higher than
ammonia [23,24,27,28]. Hydrogen has a faster rate of combustion and a higher level of heat
generation compared to ammonia and other fuels. Thus, it can be used in conventional
combustion systems with a few changes to the combustion system (engines, furnaces or
gas turbines) design [18].
Hydrogen energy can be optimized through electrochemical processes such as fuel cells
and hybrid renewable energy systems to obtain the most economical low cost energy [29–31].

(a) (b)

Figure 4. Hydrogen properties, (a) volumetric and gravimetric density of hydrogen, (b) specific
energy density compared to other fuels. Reprinted with permission [27,28].
Energies 2022, 15, 8246 6 of 49

2.2. Hydrogen Production Sources and Pathways


Hydrogen can be produced from renewable or non-renewable natural feedstocks and
energy sources through several pathways of various types, as shown in Figure 5. Green
hydrogen is obtained from renewable feedstocks and sustainable energy sources, such as
water, wind and solar power, without any harmful gases during production processes and
use. Moreover, green hydrogen can be a sustainable and secure energy, and the required
power for this process is generated by clean electricity. However, this type of hydrogen is
the ultimate goal of climate-promoting but yet is not practical [15,32,33]. Blue hydrogen is
produced from non-renewable resources (such as fossil fuels, hydrocarbons, and nuclear
sources) and renewables. It is a low-carbon hydrogen and needs a process by which carbon
is captured and stored [3,15,16,22,33]. Gray hydrogen is produced from coal, heavy oil, and
naphtha feedstocks with high carbon-to-hydrogen ratios by a conventional gasification
process that requires pure oxygen or steam to react with the feedstocks and produce a
mixture of CO and hydrogen in the ratio range of 1.6–1.8 at a high temperature [34].

(a) (b)
(a) (b)
Figure 5. Different hydrogen (a) production pathways and (b) the estimated cost of blue, grey and
green hydrogen production. Reprinted with permission [10,32,33].

2.3. Hydrogen Production Technology


Hydrogen can be produced through a variety of technologies, such as thermochemical,
radiochemical, electrochemical, photochemical, and biochemical, and integrated systems,
such as electrothermochemical processes. Figure 6 represents different hydrogen produc-
tion technologies in terms of the energy sources [3,15,26,30,34,35].

Figure 6. Diverse hydrogen production technologies and their energy sources. Reprinted with
permission [35].
Energies 2022, 15, 8246 7 of 49

2.3.1. Thermochemical Process


Thermochemical technologies, are the main form of commercial hydrogen produc-
tion, involve high-temperature processes, including reforming (steam, partial oxidation,
autothermal, plasma and water phase) and pyrolysis. They need additional hydrogen
separation and purification units to provide a high purity of hydrogen [12,17]. As shown
in Figure 7, the main resources of commercial hydrogen production are 48% from natural
gas, 30% from petroleum reforming, 18% from coal gasification and 4% from electroly-
sis and renewable energy sources [36,37]. In addition to biomass gasification and solar
thermochemical hydrogen production, it can be produced as a by-product of some other
hydrocarbon processes, such as steam cracking of natural gas liquids to obtain light olefins
and other co-products, which can produce 3.5 Mt/year of H2 . However, most of them are
low energy efficiency, expensive, and require carbon capture [26,38].

Figure 7. Chart of hydrogen production from different resources. Reprinted with permission [37].

Thermochemical technologies for hydrogen production can be classified based on


efficiency, temperature, and feedstock requirement, as shown in Table 3 [34,39,40]. The
three methods of hydrogen production are including, reacting natural gas with steam at
high pressure and temperature, such as steam methane reforming (SMR), reacting natural
gas with oxygen and steam, such as auto-thermal reforming (ATR), and reacting natural
gas with quantities of air and pure oxygen less than the stoichiometric amounts, such as

partial ✓ of methane (POM) technologies. A comparison of their efficiencies and
oxidation
reaction temperatures shows that POM has the lowest efficiency compared to ATR, and

SMR has the highest efficiency. Accordingly, SMR is the commonly accepted method of
generating hydrogen [34,39,40].
✓ ✓
Table 3. Comparison of three methods of hydrogen production from natural gas. Reprinted with
permission [34].

Material and Energy Requirements H2 /CO Ratio


Efficiency % Temperature (◦ C)
External Heat Catalysts Pure Oxidation
Steam methane
X X × 70–85 800–1100 1.2–2.8
reforming (SMR)
Partial oxidation of
× × X 55–75 950–1500 2
methane (POM)
Auto-thermal
× X X 60–75 700–1000 1.9–2.6
reforming (ATR)
Energies 2022, 15, 8246 8 of 49

2.3.2. Electrochemical and Photolytic Process


Electrochemical methods that use electricity to produce hydrogen at near-ambient
temperatures and pressures without any requirement of a purification system, such as water
electrolysis, have high conversion rates. Water electrolysis is one of the main technologies
for generating highly pure hydrogen that can provide a sustainable clean hydrogen process.
Producing 1 ton of hydrogen requires 9 tons of continuous supply of pretreated water with
about 50 MWh of electric power for operation. Meanwhile, the electrolyzer, operated most
often in base load mode, has significant value to add by participating in electricity markets
and thus minimizing electricity costs [12,17,41,42].
The photocatalytic process of decomposition uses radiation energy for decomposition
at room temperature in photo-electrolysis, photocatalysis, and photo-biological processes.
However, this technology is still in the early stages of development, and material costs and
practical problems have not been resolved [2,35,43,44]. The electrochemical and photolytic
processes of ammonia decomposition are explained in more detail in the related part of
Section 6.
Alternatively, high-purity hydrogen can be produced as a by-product of mature
industrial technologies, such as alkaline and proton exchange membrane (PEM) and low-
temperature electrolysis (LTE) from renewable electricity, on a large scale, with high energy
efficiency (PEM electrolyzer 72%) and a fast response rate. The chlor-alkali (CA) process,
which produces sodium hydroxide and chlorine through electrolysis of a sodium chloride
solution, produces a by-product of high-purity hydrogen (>99.9%) on a large scale (up to
0.4 Mt/year) without additional purification processes at a price of about USD1/kg H2 ,
which is relatively cheap. High-temperature electrolysis using a solid oxide electrolysis
cell (HTSOEC) produces hydrogen with higher electrical energy efficiency than LTE due
to the requirement of heat rather than electricity for processing energy [26]. Hydrogen
production from renewable energy sources such as water electrolysis, the most feasible
technology, is shifting from alkaline electrolysis to the more flexible PEMs and being scaled
up in the EU, the US, Japan and China; however, PEMs have the lowest energy and exergy
efficiencies [26,34,45].
Hydrogen, electricity, ammonia and heating energy can be produced simultaneously
through the combination of a solar cell electrolyzer, solar–steam cycle and wind-powered
electrolysis process. For example, the production of 10 MW of electricity, 994.5 tons/year
of hydrogen and 7201 tons/year of ammonia with a maximum energy efficiency of 50%
was reported in large cities in China and India [46–48]. Moreover, hydrogen production
using nuclear power is carbon-free, but it has more environmental, health and safety issues
than other fossil fuels [25].

2.3.3. Biological Technologies


One way of achieving very clean large-scale hydrogen production is by using renew-
able biomass sources. The process requires the preparation of feed stocks such as wood,
grass, agricultural products, crop residues, plant and animal wastes, municipal solid wastes,
food scraps and algae by several methods, such as gasification, pyrolysis, supercritical
water gasification, liquefaction, anthropogenic extraction of hydrogen from waste materials,
and biochemical processes of microorganisms, such as combined dark fermentation and
anaerobic digestion. This process needs more development, and its conversion efficiencies
are low [2,25,35].

2.4. Energy and Exergy Analysis of Hydrogen Production from Ammonia


Analyzing energy and exergy efficiencies is an important tool for optimization and
improvement, as well as the design of hydrogen production systems through the evalua-
tion and prediction of thermodynamic process defects. The exergy of ammonia cracking
describes a measure for identifying and explaining the benefits of sustainable energy and
technologies, and the exergy of ammonia cracking clearly identifies efficiency improve-
ments and reductions in thermodynamic losses for more sustainable technology. Therefore,
Energies 2022, 15, 8246 9 of 49

the exergy efficiency of an ammonia cracking system is related to its economics, environ-
mental impact, and sustainability [49,50].
The performance of a hydrogen production system based on efficiency as the ratio
of the energy or exergy content of the hydrogen to that of the material and/or energy
resource is obtained from the following equations, which indicate that when the system
irreversibility or the exergy loss is minimized, the exergy efficiency can improve and reach
the maximum value [50].
. .
m LHVH2 Eout
(Energy efficiency) η % = . = .
Ein Ein
. ch .
m exH2 Exout
(Exergy efficiency) ψ = . = .
Exin Exin
.
where η is the energy efficiency, m is the mass flow rate of the produced hydrogen, LHVH2
.
is the lower heating value of hydrogen (121 MJ/kg), Ein is the rate of energy input to the
.
process, and Eout is the total outlet energy transfer rate. ψ is the exergy efficiency, exch
H2 is
. .
the chemical exergy of hydrogen, Exout is output or useful exergy, and Exin is the exergy
input rate supplied to the hydrogen production process [50,51].
The efficiency of different hydrogen production methods depends on their emissions,
total cost, and energy and exergy efficiencies. It can be categorized in the range of 0 to 100,
as shown in Table 4, where 0 means poor performance assigned to the cost and emissions,
and the highest value of 100 means higher efficiencies, which indicates the ideal case with
the lowest costs and zero emissions [1,50,51]. Hydrogen production by fossil fuel reform-
ing, plasma arc decomposition, and coal gasification and biomass gasification methods
have high energy and exergy efficiencies. Biomass-based hydrogen production methods
have significantly low emissions, but they have low system efficiencies, significantly high
acidification potential (AP), and relatively high global warming potential (GWP), social
cost of carbon (SCC) and production costs, as demonstrated in Table 4. The ideal method
is the production of green hydrogen as it has zero emissions, the lowest cost, and 100%
efficiency [34,50,51]. Table 4 shows comparison of maximum energy and exergy efficiencies
of hydrogen production methods, which has the highest energy efficiency of 83% using
fossil fuel reforming (FFR), an exergy efficiency of 60% for biomass gasification (BIG), and
the lowest energy and exergy efficiency (2%, 1%) for photocatalysis (PCT) methods [50,52].

2.5. The Cost and Environmental Emissions


Hydrogen production from different technologies and pathways is varied in terms
of cost and environmental emissions due to economic factors, such as the cost of natural
gas and electricity. The hydrogen production cost is depending on industrial scales, the
availability of existing infrastructure, production technology, the variety of feedstocks, and
power sources. It is conventionally generated from fossil fuels and is the most cost-effective
process for large scales (above 1000 m3 /h) but releases a high amount of greenhouse gas
(GHG) (see Table 5). As the global demand for hydrogen rises, there is a need to decrease
GHG emissions by switching to clean energy and green hydrogen generated from clean
and sustainable energy sources through efficiency and industrial decarburization, but this
currently comes at relatively high energy costs [2–4,15,16,22,26,53–62].
Energies 2022, 15, 8246 10 of 49

Table 4. Comparison of the energy and exergy efficiencies of some hydrogen production methods.
Reprinted with permission [1,50,51].

Global
Acidification
Energy Efficiency Exergy Efficiency Warming
Method Cost Potential
% % Potential
(AP)
(GWP)
ELC Electrolysis 53 25 73.4 33.3 88.6
PAD Plasma arc decomposition 70 32 91.8 8.3 51.4
TLY Thermolysis 50 40 61.2 75.0 74.3
TWS Thermochemical water splitting 42 30 80.6 91.7 94.3
BIM Biomass conversion 56 45 81.0 66.7 20.0
BIG Biomass gasification 65 60 82.5 58.3 0.00
BIR Biomass reforming 39 28 79.3 62.5 8.6
PVE Photovoltaic (PV) electrolysis 12.4 7 45.0 75.0 77.1
PCT Photocatalysts 2 1 51.9 95.8 97.1
PEC Photoelectrochemical (PEC) 7 1.5 0.00 95.8 97.1
DAF Dark fermentation 13 11 75.2 95.8 97.1
HTE High-temperature electrolysis 29 26 55.4 79.2 85.7
HYC Hybrid thermochemical cycles 53 48 74.1 94.3 90.2
COG Coal gasification 63 46 91.1 0.00 13.1
FFR Fossil fuel reforming 83 46 92.8 25.0 57.1
BIP Bio photolysis 14 13 72.7 75.0 97.1
PHF Photo fermentation 15 14 76.1 95.8 97.1
APS Artificial photosynthesis 9 8 75.4 95.8 97.1
PEL Photo electrolysis 7.8 3.4 70.9 83.3 97.1
Ideal Zero emissions and cost, efficient 100 100 100 100 100

Table 5. Hydrogen production pathways and technologies. Reprinted with permission [54].

Hydrogen
Source of Energy Technology Emissions Advantages Disadvantages
Pathways
Minimal cost Considerable CO2
Huge-scale emissions
Reforming and
Grey hydrogen Fossil energies CO2 emissions production Carbon tax
gasification
Innovative Finite natural
knowledge resources
The CCUS
Economical
technology needs to
compared with
be improved
Reforming and green hydrogen
Captured carbon Investment
Blue hydrogen Fossil energies gasification and Scaling potential
emissions difficulties could
CCUS with modification
occur as green
Emitting low CO2
hydrogen becomes
emissions
more competitive
If the installation is
Significant costs
spread, it can be
Based on the cost of
more adaptable
power and the
There are no
Green hydrogen RES Electrolysis No emissions availability of water
unlimited
Minimal capacity
resources
aspects linked with
required—just
RES
water and power

Around 6% of global natural gas and 2% of coal consumption is used for producing
hydrogen, which generates about 830 MtCO2 of annual CO2 emissions and is a significant
driver of climate change [11,54]. In 2020, 60% of the annual global hydrogen was produced
from 240 billion cubic meters of natural gas, and 19% was produced from 86 billion
Energies 2022, 15, 8246 11 of 49

cubic meters of coal in China [8]. According to Figure 8, a total 90 Mt of the global
hydrogen demand in 2020 was derived from fossil fuels, 79% of which was obtained from
dedicated hydrogen production plants, and the remaining 21% was obtained from carbon
by-products. Producing 90 Mt of hydrogen can release about 900 Mt of CO2 (2.5% of global
CO2 emissions) per year. However, using CCUS technology decreased the carbon emissions
by 59.7% [3,36,55,56].

(a) (b)

Figure 8. (a) Hydrogen production methods by their source-based emissions and capture of CO2
emissions from combustion and chemical conversion, (b) the source of hydrogen produced in Europe,
2019. Reprinted with permission [55].

The hydrogen generated from green ammonia cracking demonstrates high CO2 re-
duction ranging from 78% to 95% in kgCO2 /kgH2 compared to the hydrogen production
from SMR, which is mainly gray hydrogen (96%) [48,57]. Therefore, this is regarded
as the decarburization plan most likely to achieve the goal of zero carbon emissions by
2050 [6,58,59].
Hydrogen production requires additional development to be sustainable and energy
efficient in production, usage and transportation for commercialization. However, hydro-
gen also can be produced through a large, centralized fossil fuel-based technology along
with CO2 capture and storage (CCUS), which has not yet been demonstrated technically or
commercially, and further research is needed on gas separation and hydrogen purification
processes [10,54].
In the Net Zero Scenario, the carbon emissions of energy systems can be decreased
through mitigation measures includes the direct on-site renewable energy use, improved
technology performance, energy efficiency, and behavioral changes such as energy service
demand changes from user decisions and technology developments. The technologies to
be developed include activation of grid decarburization, replacement of technologies with
electrification (heat pumps, mechanical vapor recompression), transition to renewables
sources (hydropower, geothermal, solar power, wind, marine energy), and transition to
hydrogen and hydrogen-based fuels by developing CCUS of relevant intensity [4,10,60].
The environmental impact of water electrolysis depends on the primary energy source
that provides the electricity. It cannot be an environmentally friendly technology under the
current electricity structure. It will have greater potential in the trend of power structure
transformation. However, it can be solved by using power sources from wind, solar, nuclear
and biomass. Table 6 and Figure 9 show the hydrogen production cost and environmental
impacts of different production technologies, the variety of feedstocks, power sources and
emissions [62].
Energies 2022, 15, 8246 12 of 49

Table 6. Various hydrogen production processes and their sources (feedstocks and energy) and cost
per kg of hydrogen. Reprinted with permission [28,57].

Cost
Process Sources
(USD/kgH2 )
1 SMR with CCS fossil fuels + NG 2.27
2 SMR without CCS fossil fuels + NG 2.08
3 CG with CCS fossil fuels coal 1.63
4 CG without CCS fossil fuels coal 1.34
5 ATR with CCS fossil fuels NG 1.48
6 Methane pyrolysis fossil fuels NG 1.59–1.70
7 Biomass pyrolysis Steam + wood 1.25–2.20
8 Biomass gasification Steam + wood 1.77–2.05
9 Direct bio photolysis Algae + water 2.13
10 Indirect bio photolysis Algae + water 1.42
11 Dark fermentation Organic biomass 2.57
12 Photo fermentation Organic biomass 2.83
13 Solar PV electrolysis Solar + water 5.78–23.27
14 Solar thermal electrolysis Solar + water 5.10–10.49
15 Wind electrolysis Wind + water 5.89–6.03
16 Nuclear electrolysis Nuclear + water 4.15–7.00
17 Nuclear thermolysis Nuclear + water 2.17–2.63
18 Solar thermolysis Solar + water 7.98–8.40
19 Photo electrolysis Solar + water 10.36

Figure 9. The comparison of various methods of hydrogen production in terms of cost (USD/kg) and
their environmental impacts (kg CO2 eq). Reprinted with permission [28,57].

The theoretical ideal hydrogen production option has zero production cost, no harmful
emissions, and zero social cost of carbon. The cost of producing hydrogen is highly
influenced by the scale of the installation. A techno-economic assessment showed that the
cost of hydrogen (kgH2 ) at small scales (i.e., 10 kW) reduced from 7.03 to 3.98 USD/kgH2
compared to larger industrial scales (i.e., 10 MW), and it can be decreased up to 50% more
based on sensitivity analyses [48].
The cost of hydrogen from ammonia to a commercial scale resulted in approximately
4 USD/kgH2 at 10 MW. In terms of facility costs, the PSA-adsorbent cost was 56.1%,
followed by reactor catalyst cost of 21.3%, and ammonia procurement accounted for 90.5%
of operating costs. The CO2 emission is the lowest, although the production cost is higher
compared to the existing commercial hydrogen production system, so the price difference
will be realized through the greenhouse gas emission trading system [48].
Energies 2022, 15, 8246 13 of 49

It will be critical to decrease the cost of clean hydrogen production by developing and
expanding current technologies and management of CO2 . This can significantly improve air
quality through carbon-free methods or combining them with carbon capture, utilization
and storage (CCUS). In addition, to increase the carbon tax in proportion by the amount of
generated CO2 [8].
Compared to the existing diesel energy source, a fuel-cell power system using hydro-
gen and ammonia is sufficiently replaceable in terms of weight, volume and power but
not in terms of the economy [11]. However, the electrolyzers that operate most often in
base load mode can provide significant value from participating in electricity markets by
minimizing electricity costs [8]. Historically, the cost of green hydrogen has been more
expensive than gray hydrogen from fossil fuels. However, nowadays, the cost of gray
hydrogen produced from natural gas is currently between 5.5 and 6.89 USD/kg in parts of
Europe, the Middle East, India, South Korea and Africa, while the price of green hydrogen
is between 4.84 and 6.68 USD/kg. The cost is also related to factors such as fuel price, and
transportation which mainly varies depending on geographical and economic conditions.
Policies for the use of hydrogen energy in South Korea, is shifting from current large-scale
production of gray hydrogen of 220 kt/yr, and is expect increase to production of green
hydrogen of 1.39 Mt/yr by 2030, and increase to 4.52 Mt/yr from renewable sources by
2040, while decrease in costs to KRW 3000 or 2.33 USD/kg by 2040 [15,54,63].
The unit cost of green hydrogen production with renewable electricity sources is in the
range of 2.2~8.64 USD/kgH2 ; and it could decrease due to costs of renewable sources and
the scaling-up of industry. However, it could compete with existing hydrogen production
from fossil fuel, the oil refining and ammonia production processes.
The cost of transportation is related to factors such as fuel price, which mainly varies
depending on geographical and economic conditions.
The efficiency and production volume of a hydrogen and ammonia production sys-
tem can increase using a combination of a biomass gasifier with solar heat and a power
generation system. A comparison of the hydrogen production costs based on feedstocks is
presented in Table 7 [38,62]. For example, hydrogen produced from the renewable energy
of solar heat, in a simultaneous production process of hydrogen, ammonia, urea, and
power, has an energy efficiency of 66.12%. Hydrogen prices decreased (1.94 USD/kg)
as the amount of solar irradiation increased (650 W/m2 ), and the unit cost of electricity
production was calculated as 0.084 USD/kWh [38,62,64].

Table 7. Comparing the cost of hydrogen production methods based on power sources. Reprinted
with permission [62].

Hydrogen
Hydrogen
Method Electricity Source Cost
Production (kg/day)
(USD/kg)
Wind 1400–62,950 5.10–23.37
Water electrolysis Solar PV 1356 10.49
Solar Thermal 1000 7.00
Thermochemical
Solar 6000 7.98–8.40
water splitting
Water electrolysis Nuclear 1000 4.15
Thermochemical
Nuclear 7000–800,000 2.17–2.63
water splitting
Natural gas steam With carbon capture storage - 2.27
reforming Without carbon capture storage - 2.08
With carbon capture storage - 1.63
Coal gasification
Without carbon capture storage - 1.34
Biomass gasification - - 1.77–2.05

Thermochemical cycles based on nuclear energy are low-cost and have low environ-
mental impact. Electrolysis and thermolysis with nuclear energy are more competitive in
Energies 2022, 15, 8246 14 of 49

terms of cost than with other energy sources. It is worthwhile to examine the technological
maturity and safety of nuclear energy and expect it to become the main driver for hydrogen
production in the future [62].

3. Ammonia Source, Production Pathways, Technology, and Energy Efficiency


As an ideal hydrogen carrier, ammonia is an important industrial chemical and a major
contributor to a carbon-free economy. It has a well-established infrastructure, a global trade
and distribution network, and accounts for the largest production and market of hydrogen.
There is no alternative to hydrogen in the ammonia fertilizer industry. Ammonia is being
produced in huge quantities of around 200 million metric tons annually, 80% of which is
used for producing fertilizers [65,66].

3.1. Ammonia Properties


Ammonia contains no greenhouse gases, consisting of hydrogen (17.8 wt%) and
nitrogen, with a density of 0.769 kg/m3 and a boiling point and freezing point of − − 33.3 and
− 77.65 C (Table 2) at standard pressure, respectively. Ammonia has a narrow flammability
− ◦

limit between 15 and 25% in air and a higher energy density of 22.5 MJ/kg compared to
fossil fuels (see Figure 10). It easily becomes liquid at 10 bar, to easily transport and safer
handling [12,17,62–69].

Figure 10. Comparison of the energy density of ammonia and hydrogen with carbon-based fuels.
Reprinted with permission [69].

3.2. Ammonia Production Pathways and Technologies


Ammonia is produced from different technologies, such as thermochemical, elec-
trochemical, and biological processes, in a variety of pathways, such as carbon-based or
renewable feedstocks, as shown in Figure 11 [70,71]. The Haber–Bosch process is the most
common, well-established, economic and sustainable method of ammonia production. Am-
monia production consumes a high amount of energy, about 1.8% of annual global energy,
to provide the high temperature and pressure required for the process. This process mainly
use over 30 GJ/tNH3 and produces about 2.16 kg CO2 eq/kgNH3 [39]. More than 90% of
the total global ammonia (about 176 Mt/year) is produced through SMR [72], emitting
more than 380 million tons of carbon dioxide per year, which is the highest greenhouse gas
emissions in the chemical production system, as shown in Figure 12 [42,58].
Energies 2022, 15, 8246 15 of 49

Figure 11. The pathways and technology roadmap of ammonia production in different scenarios
(the stated policies, the sustainable development, and the Net Zero emissions) from 2020 to 2050
(electrolysis pyrolysis, biomass, coal, coal with CCS, gas, gas with CCS, oil, fossil with CCU methods
and near-zero emission). Reprinted with permission [70,71].

(a) (b)

Figure 12. Comparison of greenhouse gas emissions from different liquid ammonia production
methods. (a) kg of CO2 per kg of NH3 , and (b) g of CO2 per equal MJ of NH3 . Reprinted with
permission [56].

Gray ammonia synthesized through thermochemical processes, including gasification


of coal, reforming of natural gas, naphtha reforming, pyrolysis and auto-thermal reactor
(ATR), has the characteristic of high carbon emissions [58,59,72]. Blue ammonia obtained
from the lower-carbon process is based on SMR, such as reforming and gasification of
fossil fuels combined with the CCUS process, which is an alternative source of low-carbon
hydrogen [58,59,72]. The main technologies, such as electrolysis and SMR, with or without
carbon capture while producing hydrogen, provide a low-cost energy source for mobile
applications, industrial applications, or gas grid injection [35,57,73,74].
Green ammonia can be produced from renewable feedstocks such as air and water
with sustainable energy sources using water electrolysis instead of natural gas, oil or coal,
and the energy could be relatively easy to transport and handle [58,59,67,72].
Figure 13 shows the energy consumption of various ammonia production technologies
compared to their CO2 emissions (tCO2 /tNH3 ), as well as a diagram of the Haber–Bosch
process [75]. The combination of water electrolysis with renewable feedstock technology can
be adopted for producing ammonia from renewable sources such as low-carbon hydrogen.
The price of electricity is an important factor and can be decreased by using green electricity
from renewable sources. However, it requires about nine tones of continuous supply of
Energies 2022, 15, 8246 16 of 49

pretreated water with high purity levels to produce 1 ton of hydrogen; thus, 233.6 Mt/year
of water is required for 176 MtNH3 /year [42,67].

Figure 13. Comparison of energy consumption (GJ/t NH3 ) of various ammonia production technolo-
gies and their CO2 emissions (t CO2 /t NH3 ). Reprinted with permission [75].

Figure 14 shows the comparison of various ammonia production costs (different color
types of gray, blue and green), as well as the relative CO2 emission reduction potentials
of each technology by 2030 [76]. The cost of ammonia production varies significantly by
synthesis technology as well as the regions due to price fluctuations in fuel, feedstocks
and energy, and the cost of the plant. For electrochemical methods, it is mainly related
to electricity costs (85%). The electricity cost for ammonia production without carbon
capture is between approximately 0.01 and 4 USD/kWh, which is cheaper than electrolysis
(0.015–5.0 USD/kWh) in most parts of the world, but the lowest cost regions are using
renewable solar energy with high global horizontal irradiance or onshore wind [5,7,9,71].

Figure 14. Comparison costs of different ammonia production pathways, inc


Figure 14. Comparison costs of different ammonia production pathways, including blue and green
ammonia, based on the cumulative emission savings (Kt CO2 /t NH3 ). Reprinted with permission [76].

3.3. Energy Efficiency and Exergy Efficiency of Ammonia Production


The most suitable utility systems can be defined in terms of their energy and exergy
efficiency and require minimal energy and low operating costs. According to Figure 15,
Energies 2022, 15, 8246 17 of 49

ammonia is produced from carbon-based and biomass-based feedstocks and energy sources
in various methods. The production methods can be categorized as: conventional (SMR,
at 700–800 ◦ C and ATR, at 1000 ◦ C); using electricity power, and the Rankine cycle, that
provides the required heat and power demands without gaseous fuels (WF-RC-EE: without
fuel, by Rankin cycle, and electricity); only natural gas is consumed as fuel (NG-RC-no
EE: natural gas, Rankin cycle, no electricity); natural gas and a combined cycle supplies
the required heat and power (NG-CC-no EE: natural gas, combined cycle, no electricity);
syngas is consumed as fuel for heat and power (SG-RC-no EE: syngas fuel, Rankin cycle,
no electricity); and SG-CC-no EE (syngas fuel, combined cycle, no electricity) [69].

Figure 15. Comparing exergy consumption for the various methods of ammonia production, includ-
ing conventional, WF-RC-EE, NG-RC-no EE, NG-CC-no EE, SG-RC-no EE, SG-CC-no EE. Reprinted
with permission [69].

The exergy efficiency of ammonia production from natural gas is 65.8% and from biomass
plants is 41.3%, with an average overall emission range between 0.5 and 2.3 tCO2 /tNH3 .
Biomass gasification has lower efficiency but is cheaper than natural gas, and there exists the
possibility of using the produced syngas for electricity generation [69]. The exergy consump-
tion (GJ/tNH3 ) for the conventional system reaches 32.34 GJ/tNH3 , with environmental CO2
emissions of 1.75 t CO2 /t NH3 due to the use of natural gas (Figure 16) [69]. The exergy losses
of ammonia production can be reduced by improvements in the waste heat, changes in the
system design of the process and improvements in the performance of the unit operations. As
shown in Figure 17, the exergy losses of the gasifier unit were 38.4% for SG-RC-no EE and
46.7% for WF-RC-EE [69,77–79].

Figure 16. Comparing CO2 emissions for the various methods ammonia production, including
conventional, WF-RC-EE, NG-RC-no EE, NG-CC-no EE, SG-RC-no EE, SG-CC-no EE. Reprinted with
permission [70]. The CO2 balance as a measure of the total amount of carbon dioxide emissions that
are directly or indirectly caused by production activities [69].
Energies 2022, 15, 8246 18 of 49

Figure 17. Exergy loss for the important components of ammonia production systems, such as reactor,
compressor, and reformer for conventional, conventional, WF-RC-EE, NG-RC-no EE, NG-CC-no EE,
SG-RC-no EE, SG-CC-no EE. Reprinted with permission [69].

Sustainable ammonia production can be achieved by using renewable energy sources


instead of carbon-based fuels in low-cost regions or transporting them from places where it
is relatively cheap. However, the current cost of producing green ammonia is still higher
than blue ammonia [5,9,12,17,42].

4. Development of Technologies for Converting Ammonia to Hydrogen


Regarding natural renewable storage, hydrogen and ammonia are both chemical
energy carriers and potential fuel sources, which their energy is released with no carbon
emissions when used as a fuel, providing an optimal method for sustainable energy. The
energy density by weight of hydrogen is 120 MJ/kg, which is approximately six times
higher than ammonia; hence, it has gained popularity as an alternative fuel. Ammonia is
considered for indirect storage of hydrogen because of its energy density of 22.5 MJ/kg
at HHV, approximately half of that of gasoline (46 MJ/kg) and ten times more than metal
hydrides in batteries [34,62,80].
There are several technologies for converting ammonia to hydrogen, such as thermo-
chemical, photocatalytic, and electrochemical methods [12,17,43,44,81].

4.1. Thermal and Thermo-Chemical Cracking


Thermal cracking is a typical method of hydrogen generation from ammonia. The first
thermal cracking of ammonia was conducted by Burke in 1933, where liquid ammonia was
heated in the temperature range of 550 to 650 ◦ C and flowed through tubes over a suitable
catalyst, and of 90% the ammonia decomposed [18,67,82,83].
Hydrogen can produce forms of ammonia using plasma electrolysis technology, that
can be an alternative to thermal catalysts for instantaneous ammonia decomposition in
less than a microsecond to satisfy the requirements of engines [84]. There are a variety of
plasma decomposition reactors, such as thermal plasma, cold plasma, dielectric barrier
discharge, micro-hollow cathode discharge, and warm plasma. The RF thermal plasma
reactor of 13 kW produced hydrogen at a flow rate of 28 L/min and a conversion rate of
98% but with a low energy efficiency [84]. High-purity hydrogen was produced in cold
plasma with low generation rate of 20 mL/min because of the long residence time in the
discharge zone. Large-scale real-time hydrogen production using a low-temperature arc
and
plasma had a higher energy efficiency (783.4 L/kWh (liters per kilowatt hours)) than other
plasma methods and increased to 1080.0 L/kWh after adding the NiO/Al2 O3 catalyst but
produced low-purity hydrogen (<34.8%) [84]. Warm plasma generated by a non-thermal
arc plasma can generate enough reactive species and maintain moderate gas to decompose
ammonia at temperatures between 1500 and 4000 K [84].
Energies 2022, 15, 8246 19 of 49

4.2. Thermodynamic Concerns of Ammonia Cracking


Ammonia cracking is the reverse of its synthesis reaction. It requires an energy input in
a mildly endothermic reaction at a relatively high temperature under standard pressure [85].
The enthalpy changes of NH3 are given with the following equation:

NH3 → 1.5 H2 + 0.5N2 , ∆H298 = 46 KJ/molNH3

Practically, this conversion depends on the reaction temperature and catalyst and can
be performed at higher temperatures (823 to 1023 K). Thermodynamically, the decomposi-
tion of ammonia to hydrogen is possible at temperatures of 698 K, but the decomposition
reaction is very slow and using catalysts can help to reach a higher conversion of ammonia
at operating temperatures lower than 723 K [85]. However, different from the equilibrium
condition, ammonia does not fully decompose into hydrogen, and the conversion rate is
less than 100% [35,53]. Therefore, maximum single-pass conversions require relatively high
temperatures in the range of 250–700 ◦ C, which are theoretically obtainable. Table 8 repre-
sents the equilibrium conversion values of ammonia to hydrogen at different temperatures
under normal atmospheric pressure. The highest ammonia conversion (>99%) happened at
temperatures above 400 ◦ C under normal pressure, after which the reaction was considered
irreversible and less dependent on temperature following first-order kinetics. Thus, it may
be interpreted in terms of kinetics ratherpNHthan thermodynamic limitations [85]. Hence,
r = α( ) β = 0.5
the conversion of ammonia faces thermodynamicpH limitations and kinetic barriers, which
require higher temperatures [53]. Moreover,
α and βthe kinetics of ammonia conversion depend
on the temperature and the ammonia concentration [86].

Table 8. Equilibrium ammonia conversion depends on temperature. Reprinted with permission [85].

Temperature (◦ C) 250 300 350 400 450 500 600 700


NH3 conversion (%) 89.20 95.70 98.10 99.10 99.50 99.70 99.90 99.95

4.3. Catalytic Reaction Kinetics of Ammonia Cracking


Ammonia with high contents of hydrogen under high temperatures in the presence of
a suitable catalyst is decomposed into carbon-free hydrogen. This technology is high energy
consumption (total energy efficiency 58%), requiring reactor and catalyst development,
process integration, and economic development [18,67,82,83]. The catalytic ammonia
conversion rate varies depending on many factors, such as the temperature, pressure and
catalysts, composition and purity of the gas phase, a specific surface area, and composition
of the catalyst on which the reaction proceeds. The gas phase ammonia be adsorbed by the
catalyst surface, and decomposes to hydrogen and nitrogen (see Figure 18) [53,87–92].

(a) (b)

Figure 18. Catalytic ammonia decomposition, (a) schematic, (b) conversion with temperature.
Reprinted with permission [92].
pNH
r = α( ) β = 0.5
pH
α and β

Energies 2022, 15, 8246 20 of 49

The ammonia conversion degree (αNH3 ) can be defined based on the inlet and outlet
composition of the gas phase of hydrogen and ammonia [90]. from the following equation:

XH2 F0 − F0H2 rdecomp


αNH3 = 0
=
FNH3 (1.5 − XH2 ) F0NH3

where XH2 is the molar concentration of hydrogen in the reactor in mol·mol−1 , and F0 is
the total molar flow rate of the inlet stream in mol·s−1 , FH2 0 and FNH3 0 are the hydrogen
and ammonia molar flow rates in the inlet stream, respectively, in mol·s−1 , and rdecomp is
the reaction rate of the catalytic ammonia decomposition [90].
The rate of ammonia decomposition over the iron-based catalysts shows a very com-
plex dependence on the temperature as well as on the partial pressures of hydrogen (pH2 )
and ammonia (pNH3 ). It can be obtained from the following relation:

pNH3 2
r=α β = 0.5
pH2 3

where r is the decomposition rate, α and β are rate parameters of the Temkin–Pyzhev
equation, and pNH3 and pH2 are the partial pressures of ammonia and hydrogen, respec-
tively [86,90].
The ammonia conversion rate enhances from 0.61 to 0.99 as the reaction temperature
increases from 450 to 773 K under atmospheric pressure. High conversion of ammonia
(99.85%) (see Figure 19) can be reached at temperatures higher than 700 K [80,93]. At a
constant temperature, the conversion of ammonia decreases with increasing pressure, as
shown in Figure 19. However, at lower temperatures, it requires a catalyst [94]. The ammo-
nia conversion to hydrogen rates was mainly in the range of 2–10% at low temperatures
and 3.5% at 523 K for similar catalyst materials [53].

Figure 19. Ammonia conversion association with reaction temperature. Reprinted with p
Figure 19. Ammonia conversion association with reaction temperature. Reprinted with permission [93].

The decomposition efficiency linearly increases with the reaction temperature in a


certain range and under a given flow rate. The volumetric flow rate of ammonia fed into the
reactor is an important factor that determines the reaction temperature and decomposition
efficiency of the catalyst [94]. The slowest step in ammonia decomposition is the desorption
of nitrogen. In addition to the desorption of nitrogen, the surface reaction and hydrogen
separation, also can effect on the reaction conversion and hydrogen yield [95,96].

5. Process of Converting Ammonia to Hydrogen


Ammonia cracking is a process of producing hydrogen from ammonia decomposition
over a catalyst at high temperatures and is preferentially performed at normal pressures.
Energies 2022, 15, 8246 21 of 49

Thermal reactions start at temperatures higher than 773 K without needing a catalyst, while
most of the catalytic cracking happens in the presence of a catalyst at temperatures lower
than 698 K, with a higher efficiency of about 98–99% [48].
The general process flow diagram at scales between 10 kW and 10 MW represents
the heat and mass balance of a system (Figure 20). The process starts by releasing pure
ammonia from liquid storage at a flow rate of 221.5 kg/h (at 500 kW scale) through a feed
valve (10–2.5 bar). It is first preheated in heat exchanger (HXpr) by hot output gas of reactor
and then its temperature increases to 525 ◦ C with burner combustion (HXref) at a for the
reformer. The output gas is then separated under pressure by pressure swing adsorption to
produce pure hydrogen and finally discharged at 20 ◦ C [48].

Figure 20. Schematic process flow diagram of the conversion of ammonia to hydrogen and the
separation and purification for the simulation model at a 500 kW scale with the output of H2 .
Note: HX comp = compressor cooler; HXpr = preheater; HX ref = reformer heat exchanger mantle;
REF = reformer; C = compressor; ads = adsorber. Reprinted with permission [48].

5.1. Reactors Technology


Various reactor technologies can be used for the ammonia decomposition reaction,
such as microwave, catalytic, plasma, membrane, multistage, continuous tubular, fixed bed,
batch and flow bed (Table 9). The catalytic performance of a reactor is related to its type
and size, partial pressure, pressure drop, and operating conditions, such as the feed flow
rate of ammonia, reaction temperature gradient, and sweep flow rate [35,81,97–109].
In general, in addition to the reactor type, the flow rate of ammonia, partial pressure,
dimensions of the reactor, pressure drop, temperature gradient, and sweep gas all play
important roles in governing the overall catalyst performance and operating conditions
(e.g., reaction temperature, NH3 feed flow rate and sweep flow rate) [35].
Reactors used in thermal processes require a high-temperature heat exchanger source
in addition to the development of catalytic materials [35]. The development of reactor
technology for ammonia decomposition can improve reaction kinetics by optimizing the
reactor design and relevant conditions in addition to using catalysts to overcome the heat
and mass transfer limitations [81]. The differences in conversion can depend on gas flow
conditions, deposition methods, and precursor types [53].
Energies 2022, 15, 8246 22 of 49

Table 9. Ammonia decomposition reactors, their performance, and their reaction temperatures.
Reprinted with permission [81].

Type of Reactor Properties Limitation Performance


Large temperature gradient The higher the feed
High pressure drop; difficulty limits the extent of ammonia temperature, the higher the
Fixed Bed Reactor
in maintaining flow rates decomposition; not suitable conversion; temperature
for fast catalytic reactions range 600–900 ◦ C
Conversion 97.0% with feed of
Characteristic length in the Difficult to scale up for NH3 –H2 mixture;
order of sub mm; eliminates industrial applications; also temperature 400–700 ◦ C;
Microreactor
temperature gradient; good shows mass transfer L = 55 mm D = 16 mm;
for fast catalytic reactions limitations in some cases 10 NmL/min P = 1 bar
Efficiency = 10.4%
Consists of catalyst-covered Conversion depends on pore
Conversion: 85%; flow rate:
pillar-like structures within a size; tradeoff required
Micropost Reactor 15 NmL/min; T = 650 ◦ C
microchannel to reduce mass between conversion and pore
P = 1 bar
transfer limitation size; has not been reported yet
Use of Pd membrane reduces
Removal of H2 from the
H2 selectivity; mostly
reaction zone enhances Conversion:(55–99)%,
Fluidized Bed Membrane reported using dilute
equilibrium conversion at low 50–400 NmL/min
Reactor ammonia; partial pressure of
temperature; lowers system T = 425–500 ◦ C; P = 1–5 bar;
ammonia alters conversion;
capital and operating costs
sweep gas dilutes H2
Conversion of 99.93% was
Fixed retentate pressure of 10
attained at a temperature of
Used-metal-coated sandwich bar had to be maintained to
Catalytic Membrane Reactor 550 ◦ C (Conversion: 74.4%;40
membranes inside catalyst bed ensure sufficient permeation
NmL/min T = 450 ◦ C;
of H2
P = 1–3 bar)

5.1.1. Fixed Bed Reactor


Fixed bed reactors, as the most common catalytic reforming reactor for hydrogen
production, suffer from poor heat and mass transfer behavior, high temperature gradient,
catalyst sintering, coke deposition and dust jamming. The fixed bed reactor enhanced the
conversion by increasing the feed temperature to the range of 600–900 ◦ C [85,96,98–101].
The performance of the fixed-bed reactor for ammonia decomposition was improved
using the cobalt-based catalyst, which obtained an ammonia conversion of 87.2% and a
hydrogen production rate of 29.2 mmol H2 gcat−1 min−1 at 500 ◦ C with a space velocity
of 30,000 mL gcat−1 h−1 [102]. Furthermore, the ammonia conversion over a Ru-based
catalyst (with different Cs) (Figure 21) was improved by increasing the Cs/Ru molar ratio
up to 4.5 at lower temperatures of 325 ◦ C but did not increase over a Ni-based catalyst in a
fixed-bed reactor at 800 ◦ C [89,103].
Various types of catalytic reforming reactors can be used based on their properties
and their performances. The different types of reactor technologies with inter-stage heating
and some differential elements can be found in [99,101,105]. The optimization of reactor
parameters with a pure phase of Ni2 Mo3 N catalyst using a chelating method of preparation
for high catalytic activity resulted in an ammonia conversion rate of 97% at 525 ◦ C [103]. A
schematic diagram of the preparation of hollow fiber reactors, the deposition of Ru-NCX on
the Al2 O3 substrate through sol-gel, and pyrolysis and Ru incipient wetness impregnation
methods is displayed in Figure 22 [110]. The on-board hydrogen production technology
uses hollow fiber converters (HFC) in compact ammonia cracking reactors due to the
advantage of minimizing the weight and space of the catalyst (smaller volume of 93%,
catalyst loading of 80%) while enhancing ammonia decomposition efficiency compared to
general packed bed reactors (PBR). The HFC reactor represents a uniform distribution of
the catalyst on the substrate and is cheaper than the PBR (Figure 23). The HFC is cheaper
because it uses less precious metal-based catalyst (80 wt%). It is more efficient in mass
transfer, has lower pressure drops of 99%. It can improve the residence time distribution,
Energies 2022, 15, 8246 23 of 49

and improves the reaction kinetics (internal and external diffusion limitations), which
results in a larger conversion rate (4 times at 773 K) than PBR. Therefore, it has the potential
for economical production [110].

Figure 21. Ammonia decomposition with the tubular fixed-bed reactor and Cs/Ru catalyst. Reprinted
with permission [38,103].

Figure 22. (A) the morphology of the hollow fiber, after the sol-gel impregnation with the precursor
xerogel solution (B) SEM images of the cross-section of the 4-channelled hollow fiber, at different
magnifications (i–iv), and the use of a precursor liquid solution results in the catalyst being uniformly
dispersed. (C) deposition layers of the Al2 O3 hollow fiber with the Ru-NCX catalyst in three steps of
polymerization, carbonization and incipient wetness impregnation. Reprinted with permission [41,110].
Energies 2022, 15, 8246 24 of 49

Figure 23. (A) Compression of the fixed bed PBR and the HFR decomposition rate by temperature,
(B) the size of reactors and their pressure drop per length, (C) designed reactor for 72 m3 /h H2 supply
of a 100 kW car. Reprinted with permission [41,110].

5.1.2. Fluidized Bed Reactor


A fluidized bed shows its advantage in the industrial-scale catalytic reforming process
because the coke formed on the surface of catalysts could be separated continuously. The
features such as the sweep gas effect, bubble-to-emulsion mass transfer, densified zone
formation and concentration polarization are impacted by the fluidized bed membrane
reactor [85,110–113]. Moreover, the membrane-assisted fluidized bed reactor has good
gas–solid contact and heat and mass transfer characteristics, which can increase catalytic
reforming and improve hydrogen production [85,96,98–101].
However, conventional Pd-based fluidized bed membrane reactors (FBMR) are not
suitable for ammonia decomposition due to their low performance, under the same operat-
ing conditions. Also due to weight, volume, and start-up time, they are not applicable for
PEM fuel cells in on-board hydrogen production [81,85,96,99–101,105].

5.1.3. Membrane Reactor


Membrane reactor technology with a high conversion rate can be used for ammonia
decomposition with a high hydrogen flux. It is prepared by coating the porous supports
and protective layers, such as a porous ceramic tube with a Ni/La-Al2 O3 catalyst or a
hydrogen-selective silica membrane deposited on porous tubular alumina Ru/γ-Al γ 2 O3 /α-
α
Al2 O3 bimodal catalytic support. (Figure 24). A palladium membrane reactor (PMR)
supported by porous stainless steel with a Na/Ru-carbon catalyst can be used for ammonia
decomposition due to its kinetically enhancing effect [39,105]. The hydrogen separated
and generated with a high-purity simultaneously within the same unit [99,104–107]. The
ammonia conversion can increase the thermodynamic equilibrium and kinetic enhancement
effect of hydrogen removal through the membrane walls. High hydrogen separation
efficiency can be achieved at lower temperatures compared to conventional systems. Using
a membrane reactor combined with a solar heat absorption system can lead to energy and
economic benefits [98,99,104].
Energies 2022, 15, 8246 25 of 49

Fi
Figure 24. The effect of reaction temperature of the ammonia decomposition over a Pd-based
membrane (PMR) on hydrogen recovery and ammonia concentration in the hydrogen permeate.
Reprinted with permission [39].

The membrane reactor recovers the hydrogen of ammonia decomposition and simul-
taneously separates it with high purity at lower temperatures and lower costs compared to
conventional systems. Tables 10 and 11 demonstrate the performance of hydrogen recovery
and the permeation through the Pd membrane at different temperatures for the Cs/Ru
catalytic membrane reactor [106,107].

Table 10. Hydrogen permeating through Pd membrane before and after ammonia decomposing at
different temperatures. Reprinted with permission [36].

Permeation
350 (◦ C) 400 (◦ C) 450 (◦ C)
(mol cm−2 s−1 pa−1 )
Before
H2 4.19 × 10−11 5.52 × 10−11 6.46 × 10−11
N2 Not measurable at 5 × 105 pa
After
H2 6.64 × 10−11 8.25 × 10−11 9.91 × 10−11
N2 4.42 × 10−14 N/A N/A

− − −
Table 11. Comparison of the performance of catalytic membrane reactor with the packed bed
membrane reactor. Reprinted with permission [36].
− − −
Specifications of Reactor CMR CMR PBMR
Temperature (◦ C) 400 450 520
Pressure (Mpa) −0.5 0.5− 0.3 −
Ammonia flow rate (cm3 /min) −
61.3 207.3 150
Conversion (%) 98 95.7 98
Purity (%) 98.7 99.7 99.2
Recovery (%) 85.7 78.6 66
Productivity (mol m−3 s−1 ) 8.1 23.9 3.6
Ru loadings (mgRu/cm2 ) 1.43 1.43 11.68

The performance of a conventional reactor can be enhanced by adding a membrane


system, which leads to achieving a conversion beyond the thermodynamic restrictions.
At temperatures above 425 ◦ C, the ammonia conversion was higher than 86%, and the
hydrogen purity was 99.998%. The application of a vacuum to the permeable side of the
membrane leads to higher hydrogen recovery and enhanced ammonia conversion in com-
pare with thermodynamic equilibrium conversion. Furthermore, the ammonia conversion
was not majorly changed by the reactor feed flow rate and operating pressure [104].
The various types of membranes, such as microporous ceramic, crystalline and amor-
phous, dense metal and proton conducting, perovskite and non-perovskite, are displayed
in Figure 25 [108]. The influencing factors of membrane separation are including membrane
Energies 2022, 15, 8246 26 of 49

material, membrane sealing, membrane pollution, carbon monoxide poisoning, and mem-
Fi
brane arrangement [109]. In the case of dense metallic membranes, hydrogen adsorbed
at the surface dissociates into two hydrogen atoms that can diffuse through the metal
lattice [109].

− −

Figure 25. Membrane reactors based on their different properties. Reprinted with permission [37].

5.1.4. Other Types of Reactors


The solid oxide fuel cell (SOFC) directly generates electricity efficiently through hy-
drogen produced from ammonia at temperatures between 500 and 800 ◦ C with low energy
efficiency
− − − (Figure 26). Therefore, it is requiring to development of catalysts, for improving
conversion rates at low temperatures [111,112].
Microfabricated reactors−
and some monolithic reactors

were used for−
the decomposi-
tion of ammonia with a flow rate of 500 NmL/min over a spherical Ni–Pt/Al2 O3 catalyst
at atmospheric pressure, in portable fuel cell power supply. High hydrogen conversion
(>99%) and operability were reached at a temperature of 873 K, but their energy efficiencies
− − −
were low (<15%). The porosity and permeability of the packed bed are not parameters of

engineering importance, and hence, they did not significantly affect the performance of
the reactor [99,113]. There was comparison of microreactors characteristics between the
micropost (conversion of 85%, flow rate of 15 Nml/min, and temperature of 650 ◦ C); with
the microreactor (conversion of 97.0%, flow rate of 10 Nml/min feed of NH3 –H2 mixture,
temperature of 400–700 ◦ C) [103].
The microwave reactor has a high conversion at lower reaction temperature due to its
selective heating properties, which can directly transfer the energy required for the ammo-
nia decomposition reaction from the microwave systems to the active surface; however, the
formation of a hot zone and heat losses are problems that need to be solved [114].

Figure 26. Schematic reactor integrated with SOFCs with combined heat, power and rate of ammonia
decomposition [113,114].
Energies 2022, 15, 8246 27 of 49

5.2. Catalysts for Ammonia Cracking


Ammonia decomposes with adsorption over the catalyst surface and releases hy-
drogen in a stepwise sequence reaction. High ammonia conversion is possible at lower
temperatures thermodynamically, but yet higher reaction temperatures are required (due
to kinetic limitations), as well as the presence of an active catalyst. In this regard, differ-
ent catalysts, promoters and support materials can be used for the required industrial
application. Besides these, the supports metals and binding energy can be controlled for
catalyst efficiency and suitability. The development of catalysts for economic and high-
performance decomposition depends on their activity, long-term thermal stability, lifetime,
lower pressure drop and system integration [48,113].

5.2.1. Catalysts Characterization, Activity and Performance


Catalytic ammonia cracking is a general technology for producing hydrogen. These
catalysts consist of a transition metal, a promoter, supporter, and/or a second catalytic
metal. The several factors need to be considered in the assessing of different catalysts, such
as, pure metal catalysts, oxides or alloys, cost, long-term stability, and efficiency. This could
include noble or precious metals and non-precious metals based on their costs, types and
properties [112,115].
The most precious metals used as catalysts in catalytic decomposition include Cr, Co,
Cu, Fe, Ir, Ni, Pd, Pt, Rh, Ru, Se, and Te, alloys of aluminum oxide with nickel, Ru and Pt,
and alloys of iron with other metal oxides, and examples of non-precious metals include
Al, Ce, Si, Sr, and Zr [43,44,115,116].
Catalysts can also be categorized into perovskite and non-perovskite types based
on the similarity of their crystal structures to mineral calcium titanium oxide (CaTiO3 ).
There is a subgroup of perovskites that consists of heavier halides (Cl, Br, and I), both
fully inorganic and hybrid organic–inorganic ones, as well as the many variants. They
exhibit unique sensitivity and long-term stability. Moreover, catalysts with metal–support
interaction enhancement, the application of promoters, and bimetallic alloys have also been
developed to overcome the high activation energy barrier [103,117,118].
Catalyst reaction rates and catalytic activity can be enhanced by the type of active
metal, type of support material with good physical properties, particle size, surface area,
dispersion of the catalyst, and activity of the promoting material. The presence of the
additives and the alteration of the support material can modify the nitrogen desorption step
and the catalytic properties of the catalyst, such as Al2 O3 –TiO2 , which is used commercially
as a support for the catalysts [119,120].
The catalytic activity and performance of different single-metal catalysts are in the
order of Ru > Ni > Rh > Co > Ir > Pt ∼ = Fe> Cr > Pd > Se ∼ = Cu > Te > Pb, respectively [101].
Various high activity elements can be used as catalyst supports to replace the expensive Ru
for the economical cracking of ammonia to hydrogen [121]. The catalytic activity varies
with different ammonia concentrations and different types of support materials, such as
platinum or palladium, or for example, for Ni components, i.e., Ni/Y2 O3 > Ni/Gd2 O3 >
Ni/Sm2 O3 > Ni/La2 O3 > Ni/Al2 O3 > Ni/CeO2 [121,122].
The performance of catalysts can be evaluated by the turnover frequency (TOF) as a
quantifying rate of reaction in terms of catalytic activity and decomposition conditions. The
TOF is the number of moles of reactant consumed or moles of product per mole of catalyst
per unit of time (s–1 or h–1 ), which is significantly influenced by the catalytic activity of
different catalysts metals [113,122], the catalysts nitrogen adsorption energy [118], and
particle size [86]. The catalytic activity is generally lowered with a reduction in the catalyst
particle size, for example, in nickel-based catalysts [86]. It is also influenced by the forms
(pelletized and powder) of the supported materials [48].
Energies 2022, 15, 8246 28 of 49

Moreover, the activity for single-metal catalysts, is a function of the nitrogen binding
energies (QN(0) ) [115]. The TOF is defined in terms of turnover number (TON) and time of
reaction, as in the following equation [122,123]:

TON
TOF =
time of reaction
where the turnover number (TON) is a dimensionless value, defined as the number of
moles of reactant consumed per mole of catalyst before deactivation under given reaction
conditions [53]. The TON is an important parameter for evaluating the stability of the
catalyst and is associated with the temperature, the concentration of the substrate, the
deactivation of the catalyst and its pre-activation (zero slope), as shown in Figure 27.
The decomposition conditions significantly influence their catalytic performance and can
provide a real measure of the efficiency of a catalyst [120–123].

Figure 27. Model of active catalyst concentration and the turnover frequency (TOF). Reprinted with
permission [120].

Nanocarbon technology can contribute to the commercialization of hydrogen produc-


tion from ammonia by increasing the catalyst performance, reaction surface area, reactivity,
secured durability and expanding facilities. The catalytic ability of nanostructured electro-
catalysts usually varies with size and morphology [124].

5.2.2. Ru-Based Catalysts


Ru is the most active metal among Ru, Rh, Pt, Pd, Ni, and Fe supported on carbon-
based supports and metal oxides. Theoretical studies support the fact that an Ru surface
has optimal binding energy with a nitrogen atom, leading to the highest ammonia decom-
position activity among mono-metals (Figure 28) [98,117,118]. An Ru-based catalyst for
ammonia conversion has the highest turnover frequencies (TOF) among the single-metal
catalysts, which is presented as a function of the nitrogen binding energies (QN(0) kcal/mol),
as shown in Figure 29 [109].
Energies 2022, 15, 8246 29 of 49

Figure 28. Turnover frequency (TOF) (a) as a function of nitrogen adsorption and (b) of various
precious metal catalysts. Reprinted with permission [48].

Figure 29. Comparison of turnover frequencies (TOFs) of supported catalyst for ammonia decompo-
sition by modeling (circles, left axis) and experimenting (triangles, right axis) at 850 K. Reprinted
with permission [115].

Ru-based catalysts supported on carbon materials (Figure 30) have been proposed
as the most efficient catalysts for the ammonia cracking reaction [43,44,115,116]. Using
Ru or Cs-Ru as a catalyst, carbon powder pre-treatment solutions and catalyst deposition
conditions were obtained with a maximum ammonia conversion rate of 90% and hydrogen
generation rate of 29.8 mmol/min gcat at 673 K. Ru, when supported on carbon nanotubes
(CNTs), increases decomposition conversion, achieving an ammonia conversion level of
about 84.65%, with an H2 formation rate of 28.35 mmol/min gcat , and maintaining the
effective area of the catalyst under reaction conditions, ambient pressure and a temper-
ature of 773 K. This is due to the high dispersion of Ru particles and the inhibition of
particle growth of the catalyst, resulting in the stability of the catalyst and high catalytic
activity [43,44,115,116]. The performance of the catalysts for the ammonia decomposition
reaction can be enhanced by applying them over different supported materials, such as
ruthenium-based catalysts on carbon materials, as the most suitable supports include active
carbons, high surface area graphite carbon CNTs, and carbon nanofibers [125–127].


Energies 2022, 15, 8246 30 of 49

Figure 30. Ru-based catalysts for ammonia decomposition. Reprinted with permission [115].

Due to the properties of an Ru-NCX heterogeneous hybrid nanocomposite catalyst


containing Ru, N, and C atoms, the activity and durability of the release of hydrogen
from ammonia at temperatures ranging from 475 to 550 ◦ C were enhanced compared to
Ru. If the Ru catalyst was fixed through the activation process using the CO2 of a carbon
xerogel (Figure 31), the decomposition rate was 3.5 times higher than that of the existing
deactivated catalyst at 450 °C. Carbon support, nitrogen, and sodium elements were used
to promote the activation of the catalyst [110,124,125].

Figure 31. Development process of resorcinol–formaldehyde carbon xerogel for Ru-based catalyst.
Reprinted with permission [125].

Some Ru-based catalysts are used for ammonia cracking, such as Ru/CNTs, which
have higher catalytic activity than Ru/MgO, Ru/AC, Ru/ZrO2 , and Ru/Al2 O3 respectively
(Table 12). In addition, adding a potassium (K) promoter increased the ammonia conversion
to about 97.3% and resulted in a hydrogen formation rate of 32.6 mmol/(gcat min) at an
ammonia flow rate GHSV of 60,000 mL/(gcat h) at 450 ◦ C [115]. After the Ru/CNT was
modified with potassium nitrate, carbonate or potassium hydroxide (KOH) as the best
promoter, the rate of ammonia cracking and the rate of hydrogen evolution significantly
improved to 99.74%, and the hydrogen formation rate reached 47.88 mmol/min gcat . The
Ru-based catalysts (Ru-Na/CNT) have shown a high conversion of ammonia (100%) at a
low temperature of 673 K. The high activity of ruthenium nanoparticles was with sizes in
the range of 3 to 5 nm [86].
Energies 2022, 15, 8246 31 of 49

Table 12. Summary of ammonia decomposition catalysts based on reaction temperature, performance,
conversion and efficiency rates. Reprinted with permission [115].

Catalyst/Support Temp. (K) Conv. Eff. Rates (%)


Ru/Al2 O3 at (1 bar) 673 99.00%
Ru/Al2 O3 at (5 bar) 673 96.00%
Ru/Al2 O3 at (5 bar) 773 99.00%
Ru/Al2 O3 at (10 bar) 673 92.00%
Ru/Al2 O3 at (10 bar) 723 95.50%
Ru/Al2 O3 at (10 bar) 773 97.20%
Ru/La–Al2 O3 pellet catalyst 773 99.7%
Ru/CNT treated with KOH 773 99.74%
Ru, when supported on carbon nanotubes 84.65% with an H2 28.35
773
(CNTs), mmol/min gcat
Ru or Cs-Ru, carbon powder pre-treatment
673 90%, H2 29.8 mmol/min gcat
solutions and catalyst deposition conditions
Ni components
Ni/Al2 O3 823 98.30%
Ni-CeO2 /Al2 O3 873 99.90%
Ni/SBA-15 873 96.00%
Ni well-dispersed layers on mesoporous
880 98%
γ-Al2 O3
Na/NaNH2 800 99.20%
CsH2 PO4 796 96% for 1.48 mol H2 /gcat h
Fe–MOx Ce, Al, Si, Sr, and Zr
N/Fe/TiO2 (NFT) 298 60% with radiation
CuO on TiO2 nanotube rows (TiNTAs) 298 50.1% with radiation
converts 99% into the
Rb precursor in Al-anodized Al2 O3
873 equivalent of 60 W of
micro-reactor
hydrogen
Nitride and carbide catalysts
Carbides and nitrides of Fe, Co, Ni, Ti, V, Mn, 623–923 96–98%
and Cr
Metal amides/imides, alkali metal amides as
Below 723
sodium amide (NaNH2 ), lithium amide, 99%
above 873
Li2 NH.
Bimetallic catalysts, molybdenum, cobalt,
CoMo and traces of Co added to Fe, below 600 K 96%
Ptsn/Mgo, Pd, Cu, Ge

The hydrogen production from ammonia decomposition on a commercial 5 wt%


Ru-activated carbon catalyst with different cesium (Cs) loading at lower temperatures of
325–400 ◦ C in the fixed-bed reactor showed the Cs effects on the ammonia conversion at
a higher Cs/Ru molar ratio. Enhancing the Cs/Ru molar ratio increased the ammonia
conversion with a maximum value of 4.5 as the optimum Cs loading, with almost 100%
of the ammonia converted (with the GHSV (gas hourly space velocity) from 48,257 to
241,287 mL/(h·gcat ), at 400 ◦ C [103].
According to Figure 32, the ammonia decomposition conversion increased over the
Ru (2 wt%)/La (20 mL%)-Al2 O3 at temperatures of 350 to 550 ◦ C under different GHSVs
of 5000–30,000 mL/gcat hr and catalyst loadings. The TOF comparisons of various Ru
supported on pelletized and powder forms show, among powder forms, that KCNTs have
the highest and TIO2 has the lowest, and in pelletized forms, La Al2 O3 has the highest
value of TOF [48].
Energies 2022, 15, 8246 32 of 49

Figure 32. Catalytic activities of Ru/La Al2 O3 under different GHSVs with varying catalyst loadings
(a); TOF comparisons of various Ru supported on pelletized and powder forms (b); catalytic activities
with varying temperatures (c); TEM images of durability tests of the catalyst over 280 days (6700 h)
(d). Reprinted with permission [48].

The perovskite catalyst Ru/La–Al2 O3 pellet had a high catalytic activity of 2827 h−1
at 450 ◦ C and stability for over 6700 h at 550 ◦ C, exceeding the performance efficiency of
83.6% for producing hydrogen over 66 L/min. According to Figure 32, the simulation of
the typical ammonia hydrogen production process showed that the performance of the
pellets with the Ru(2 wt%)/La-Al2 O3 perovskite structure is 98% or higher at 550 ◦ C [48].
Although Ru-based catalysts show the highest catalytic activity in high ammonia
concentrations at the temperature of 425 to 500 ◦ C, it has the disadvantage of scarcity and
quick deactivation that impedes the large-scale application of ammonia decomposition,
and it is expensive, resulting in a high cost of ammonia decomposition [102].

5.2.3. Non-Ru-Based Catalysts


Ammonia hydrogen production technology has a clear tendency to replace expensive
Ru-based catalysts with other materials [48]. High catalytic activity of the non-Ru catalysts
can be achieved by modifying the primary catalyst component, adding promoter and sup-
port materials, and using inactive metals with further treatments of surface modifications
and alloying techniques [120,128]. Ammonia decomposition at low temperatures with
advanced catalyst that are cheaper than ruthenium, need to consider the mechanisms of
the reaction kinetics, the effects of catalyst formulation and synthesis, catalyst modification
including promoters and supports, and increasing the efficiency with nanotechnology and
using two or more promoters at the same time [81].
Various types of transition metals, including promoters and supports, have been
developed to replace precious catalysts. It is important to develop transition metal catalysts
with low cost and high stability, such as Fe, Co, and Ni, etc. Among them, cobalt has been
considered a replaceable metal of Ru due to its low cost and nitrogen adsorption [102].
Polycrystalline foils and wires made of Pd and Ir were used at low temperatures from
500 to 1190 K to reduce the mass of the catalyst, demonstrating the impact of catalyst
form on the performance of ammonia decomposition and showing an improvement in the
reaction rate. The decomposition rate of ammonia on Ir was high compared with other
metals (Ir > Rh > Pt >Pd), respectively [113].
γ
Energies 2022, 15, 8246 33 of 49

Ammonia decomposition over a series of fine powders of a Fe–MOx catalyst (M rep-


resents different metal components Ce, Al, Si, Sr, and Zr) showed that Fe–(Ce, Zr)O2 was
highest because the additive (Ce, Zr)O2 solution worked as a solid acid to enhance the
ammonia adsorption and reaction probability of Fe components at relatively low tempera-
tures. A CeO2 promotor was used to increase the catalytic activity by the surface area of the
catalyst, and Ni/Al2 O3 was used for the stability of catalysts for ammonia decomposition,
resulting in a conversion of 98.3% and hydrogen formation rate of 32.9 mmol/min gcat at
823 K [43,44,115,116].
One example of an alternative catalyst includes layers of well-dispersed Ni on meso-
porous γ-alumina. This catalyst is low cost, stable and exhibits high activity, achieving
complete conversion of pure NH3 at a temperature of 880 K, which makes this catalyst a
promising candidate for in-situ H2 generation from ammonia to feed fuel cells in vehicles
or industry [43,44,115,116].
The decomposition rate increases with the partial pressure of ammonia on surface
sites of the basal planes of metals such as tungsten, platinum, and molybdenum sulfide
(MoS1.65 ) nanocrystals at high temperatures [81,129]. The activity of ammonia decomposi-
tion improves by combining bimetallic catalysts, such as molybdenum, which has a high
nitrogen binding energy with cobalt with a low binding energy and adding traces of Co to
Fe [81].
Metal-sulfide photocatalysts are used in hydrogen production, CO2 reduction, pollu-
tant decomposition, and N2 fixation due to their suitable band-gap energy, active sites, and
outstanding optoelectronic properties [113]. Other alternative catalysts to expensive metals
are nitrides and carbides of transitional metals, such as Fe, Co, Ni, Ti, V, Mn, and Cr, which
have a high activity, as their N2 nitrogen desorption is the rate-determining step [113].
The nickel and cobalt in the form of carbon-supported cobalt catalysts can have a high
ammonia conversion rate of about 350–400 ◦ C due to weaker nitrogen binding and effective-
ness activity [114]. The La-based perovskite, such as LaMO3 (M = La-Co, La-Ni, La-Ce-Co
and La-Ce-Ni) and LaCeNiO3 catalyst, can replace Ru in the ammonia conversion due to their
physical properties, the surface area and pore-size distribution. According to Figure 33, am-
monia decomposition over perovskite-type lanthanum-based catalysts increased at different
temperatures
− in the range of 300–600 ◦ C and at a GHSV of 6000 h−1 [130].

Figure 33. Comparison performance of lanthanum-based catalysts for ammonia decomposition.


Reprinted with permission [130].

The ammonia decomposition reaction over perovskite-based polymers of a lanthanum-


based catalyst consisting of calcined nickel and cobalt showed a higher catalytic perfor-
mance for La- and Ni (LaNiO3 ) than Co(LaCoO3 ) catalyst, with a conversion rate of 99% at
450 ◦ C (Figure 34) [131].
Energies 2022, 15, 8246 34 of 49

Figure 34. The process of ammonia decomposition over a La-based catalyst consists of calcined nickel
and cobalt [131].

Nickel-based catalysts can be used for ammonia decomposition as highly efficient


non-Ru catalysts. The catalytic performance of catalysts is significantly influenced by the
properties of the support [132]. They need preparation procedures and are expensive. Their
efficiency depends on supports, which are highly active if supported on ceramic materials
and inactive on carbon materials [133]. Although, some of its components, such as the
Ni/MgAl2 O4 , can be used as an economical and highly efficient non-Ru catalyst. According
to Figure 35, the Ni/MgAl2 O4 - LDH catalyst with high thermal stability, through calcined
layered double hydroxide (LDH) supports, showed an ammonia conversion efficiency rate
of 90% at 600 ◦ C [132].

Figure 35. The effects of catalysts support properties on performance of am


Figure 35. The effects of catalysts support properties on performance of ammonia decomposition
at temperatures from 250 to 800 ◦ C (a) ammonia conversion at different temperatures (b) different
support catalysts with the highest conversion rate for the Ni/MgAl2 O4 - LDH catalyst (c) time course
of different support catalysts and (d) time course of ammonia conversion at 550 to 600 ◦ C [132].

The main obstacle is hydrogen passivation/poisoning of the catalytic surface to attain


a low-temperature conversion of the process. Furthermore, nitrogen association for this
reaction can be rate-limited over an Mo3 N2 cluster for an ammonia decomposition reaction.
Some non-precious Mo2 N-based catalysts reached the highest decomposition of 100% at
823 K [115]. Furthermore, according to Figure 36, the ammonia decomposition over Ni-
based catalysts on metal oxides supported with Ni/Y2 O3 reached 99.74%, and the hydrogen
formation rate was 47.88 mmol/min gcat as a function of temperature [111].
Energies 2022, 15, 8246 35 of 49

Figure 36. Ammonia cracking over Ni-based catalysts at different temperatures from 350 to 650 ◦ C.
the highest ammonia conversion rate at different temperatures reached with Ni/Y2 O3 . Reprinted
with permission [111].

The sizes of nickel particles have an effect on their activity and and their turnover
frequencies (TOF) in the decomposition of ammonia. As displayed in Figure 37, the nickel
particles with average sizes below 2.9 nm showed considerable activity, with an optimum
value of 2.3 nm [86].

Figure 37. Relationship between the forward ammonia turnover rate (TOF) and average particle size
(a)) for Ni (circles) in compare with Ru (triangles), (b) for Ni–Al2 O3 (solid squares) in compare with
La–Al2 O3 catalysts (hollow squares). Reprinted with permission [86].

The cobalt-based catalysts consisted of a promoter element of 1% K and the use of 5%


Co. Silicon carbide support (SiC) showed an excellent conversion performance of 100% at
450◦ C (Figure 38). The performance of cobalt was less than Ru, but it has a lower price
and thus can be used as an alternative for developing a catalyst suitable for commercial
purposes [133,134]. The Co3 O4 nanoparticles were dispersed densely on the barium hexa-
aluminate (BHA) surface, with excellent catalytic stability for high ammonia conversion
rates [102].

Figure 38. Cobalt-based catalysts, their support and promoters in ammonia decomposition. Reprinted
with permission [133].
Energies 2022, 15, 8246 36 of 49

Using alkali metal imides and amides, such as sodium amide (NaNH2 ) or lithium
amide, Li2 NH, due to their unique reaction mechanism, can increase ammonia decompo-
sition at temperatures lower than 450 ◦ C compared with the higher temperatures above
600 ◦ C required for other catalysts to reach 100% conversion efficiency [81]. A compar-
ison of the performance of lithium and sodium amide supported with nickel (Ni) and
ruthenium catalysts for ammonia decomposition into hydrogen is represented in Figure 39.
The catalytic performance of lithium for an ammonia flow rate of 500 sccm (cm3 /min) at
580–600 ◦ C was 99% (for 1 g of lithium imide), which was higher than sodium and other
catalysts [81,135].

Figure 39. Comparison of catalytic activities of metal amides with supported nickel and ruthenium
catalysts at 580–600 ◦ C. Reprinted with permission [136].

6. Other Ammonia Decomposition Technologies


This section takes a look at other ammonia decomposition technologies, such as elec-
trochemical ammonia decomposition and photodecomposition technology. The separation
processes of hydrogen stream from unconverted ammonia and nitrogen. In addition to
these, the purification of hydrogen is briefly discussed.

6.1. Electrochemical Decomposition of Ammonia


The technology of electrochemical ammonia decomposition is an alternative to high-
temperature thermal decomposition for producing high-purity and green hydrogen at
near-ambient conditions and with high conversion rates, with the potential for large-scale
development of clean energy. It includes hydrogen evolution reactions and ammonia
oxidation reactions [53,136].
The electrochemical reaction of the electrolyzer to decompose ammonia by the electric current
under ambient conditions of temperature and pressure is shown in the following reactions:

Ammonia electro-oxidation: 2NH3(aq) + 6OH− → N2(g) + 6H2 O + 6e− , E0 = −0.77 V


− → − −
Reduction of water: 6H2 O + 6e− → 3H2(g) + 6OH− , E0 = −0.83 V

Similarly, hydrogen can also→
− −
be produced from ammonia using ammonia electrolysis.
In this mechanism, an alkaline electrolytic cell is used to couple ammonia electro-oxidation
and hydrogen evolution. To date, the process has been determined to be too slow for
practical implementation. The electro-catalytic method using aqueous alkali electrolytes has
required high operating potentials, implying poor energy efficiency, and has suffered from
catalyst deactivation over time [53]. The theoretical voltage of the ammonia electrolysis
required is 95% less than for water electrolysis, for which the cell voltage is 1.223 V. However,
practically, higher energy is required to overcome the obstacle of the catalytic process
kinetics, and thus, it is important to develop efficient and highly selective catalysts [136].
Energies 2022, 15, 8246 37 of 49

Electro-catalysts in the ammonia electro-oxidation, such as Pt-based binary or ternary


alloys, can provide higher catalytic activity and stability that improve the slow kinetics at
the anode [136].
The polymer gel electrolyte reduced the onset potential for H2 evolution and conse-
quently increased H2 production, NH3 conversion and Faraday efficiency before anode
electro-catalyst poisoning occurred. The mechanical strength of the polymer gel needs to
be improved [85].
Nickel doping in cobalt-based compounds induced the active sites and consequently
enhanced electrochemical hydrogen evolution performance at lower operating tempera-
tures and produced more pure hydrogen than the thermochemical decomposition method,
resulting in enhanced electrochemical performance compared to water electrolysis [124].
Numerous promoting materials have been adopted to increase catalytic activity, in-
cluding K, Na, Li, Ce, Ba, La, and Ca. In addition, K-based compounds, such as KNO3 ,
KOH, K2 CO3 , KF, KCl, K2 SO4 and KBr, also have potential as promoting materials. These
promoting materials donate their electrons to the surface of the support material, lead-
ing to charge balance during the decomposition. The promoting material also facilitates
intermediate-step stabilization due to its low ionization energy. Moreover, support ma-
terials, which are electronically conductive, cheap and have a high surface area, are also
expected to improve catalytic activity. Potential support materials include carbon nan-
otubes, template SiO2 , porous Al2 O3 , active carbon, graphitic carbon and mesoporous
carbon [17,53,62,112,137].
The proton-conducting electrolyte cesium dihydrogen phosphate CsH2 PO4 (CDP) is a
catalyst used in electrochemical ammonia decomposition for hydrogen production without
residual ammonia problems and a performance rate of 1.48 molH2 g−1 h−1 production
at 250 ◦ C [53]. A thermal decomposition catalyst (Cs promoted Ru on nanotubes) with
an all-solid-state electrochemical conversion cell (based on CsH2 PO4 ) is a device that is
operable at 250 ◦ C. The resulting polarization curves indicate high current density at a
modest voltage (far beyond what can be attained from alkali electrolyte cells), as well as
catalyst utilization efficiency that far exceeds traditional thermal decomposition [137–139].
Hydrogen was generated in an ammonia electrochemical cell (AEC) using 0.06 V of
electricity from renewable energy on a proton-conducting membrane in the presence of a
catalyst at 250 ◦ C, which is much lower than the thermal energy required from traditional
fossil-fuel methods at 500 to 600 ◦ C. However, after removing the hydrogen, the ammonia-
splitting reaction proceeds beyond what the catalyst can achieve alone, but the generation
rate stay at 1.5 mmol/min/ per g catalyst. Performance of the systems is related to
temperature, type and concentration of impurity of hydrogen stream [140,141].
The hybrid thermal–electrochemical technology for converting ammonia to hydro-
gen integrates a solid-state proton conductor with an advanced thermal-cracking catalyst,
providing a highly efficient process free of residual ammonia without the risk of generat-
ing NOx. It directly produces highly pure hydrogen equivalent to the consumed energy
without any loss. Using a very low energy (0.4 V) results in a hydrogen production rate com-
parable to that from thermal decomposition at a high temperature of 350 to 500 ◦ C [53]. A
schematic of the hybrid thermal–electrochemical cell for on-demand ammonia-to-hydrogen
conversion is demonstrated in Figure 40. The thermal cracking layer (TCL) is adjacent to
the hydrogen electro-catalyst layer (EL), which is, in turn, adjacent to a membrane of the
solid-state proton conductor (CDP) [53]. Nanotechnologies use a CNT-Ru catalyst coated
on carbon nanotubes for the electrochemical cell. The catalytic ability of a nanostructured
electro-catalyst usually varies with size and morphology [124].
Energies 2022, 15, 8246 38 of 49

Figure 40. Schematic of ammonia-to-hydrogen conversion by the hybrid thermal–electrochemical


cell. Reprinted with permission [53].

6.2. Photocatalytic Ammonia Decomposition


The energy supply of photodecomposition technology can be provided by radiation
of an electron beam, ion beam, plasma, microwave, radio frequency (RF), and solar energy
for the decomposition of ammonia over a metal-loaded photocatalyst in a reactor to obtain
a high hydrogen yield (H2 µmol/g-cat) with the high conversion rate [113].
The microwave and μ radio frequency systems (RF) and ultraviolet radiation (UV)
technologies can supply the energy required for obtaining hydrogen from decomposing
ammonia over a photocatalyst at room temperature. The industrial microwave and RF
operate within a range of frequencies (2.45, 1.356, 0.915 GHz, and 2.712 MHz) provided by
the reaction temperature of the reactor [44,114,116].
The recent active photocatalysts, including nanostructures, C3 N4 , graphene and other
carbon-based materials, as well as their hybrid catalysts and metals with larger work
function, can provide more effective catalysts, such as ZnO and TiO2 -based catalysts
(Pt/M-TiO2 and Pt/Fe–TiO2 ) [43,44,116,142]. The hydrogen production rate increased with
increasing specific surface area in the photocatalytic ammonia decomposition over the
various platinum-based (Pt(0.1)/TiO2 ) photocatalysts (different specific surface area) in a
reactor with flowing water vapor, without flowing water vapor, and fixed-bed flow reactor,
especially for the specific surface area of less than 100 m2 /g [116,143,144].
A high rate of hydrogen production was obtained from decomposing aqueous am-
monia solutions over Pt/TiO2 photocatalysts under UV irradiation of a 500 W Xe lamp at
room temperature. The process is shown in Figure 41 [145].
The performance of ammonia decomposition when using radiant solar energy can be
optimized based on three indicators with similar optimization directions, including heat
absorption rate, entropy generation rate, and thermal efficiency. The entropy generation
rate and energy conversion rate were enhanced due to an increase in the heat absorption
rate, while the thermal efficiency decreased, as exhibited in Figure 42. Solar energy with an
intensity of 800 W/m2 , which provided a temperature of 400 K, meeting the requirements
for ammonia decomposition, resulted in an energy conversion rate of 39.6%, a thermal
efficiency of 52.1%, and an entropy generation rate of 23.9 W/K [98].
Energies 2022, 15, 8246 39 of 49

(a) (b)

Figure 41. The hydrogen production over the platinum loaded titanium oxide photocatalysts: the
highest product yield was obtained on the platinum loaded anatase. (a) variation of production rate
with a specific surface area of catalyst (b). Reprinted with permission [116,143].

Figure 42. The three-dimensional coordinate system for the decomposition of ammonia with solar
energy. The energy conversion rate shows the color of the surfaces: the lower rate, closer to blue, and
the higher rate, closer to red. Reprinted with permission [100].

The production of hydrogen from ammonia water over TIO2 catalysts (N/Fe/TiO2 )
doped with metal ions (Fe, Ag, Ni) and nitrogen using plasma energy showed an energy
efficiency and hydrogen production efficiency rate of 60%, which was higher than 50%
for conventional electrolysis, and the efficiency increased by controlling the band gap
energy [43]. The highest yield of hydrogen obtained from decomposing aqueous ammonia
solution with Fe-doped TiO2 photocatalysts (Pt/Fe–TiO2 ) is shown under UV irradiation at
room temperature [44].
Photo-electro-catalytic which achieved by depositing CuO on TiO2 nanotube rows
(TiNTAs), reaching hydrogen production from decomposition of ammonia with a maximum
efficiency of 50.1%. This reaction can be activated by the UV radiation of the mercury lamp
and control of the band gap energy at the anode by using nanostructure control and
inexpensive materials [145].

6.3. Separation and Purification Technologies


The released hydrogen stream from the ammonia decomposition reactor, has a high
purity without carbon monoxide. But it may contain some amount of unreacted ammonia,
which can damage the catalyst after recirculating through the system for hydrogen recovery.
Therefore, the output gas needs to be separated and purified [146]. Hydrogen obtained
Energies 2022, 15, 8246 40 of 49

from the water electrolysis, or from catalyst cracker system using a magnesium chloride
trap [135], has high purity and does not require separation process [66,147–149].
Generally, a hydrogen stream containing unreacted ammonia can be cooled and
compressed to liquefy the ammonia and remove it, recirculated to the plasma membrane
reactor system to be decomposed, or passed through a mineral acid solution (such as H2 SO4 )
in a series of contact scrubbers. The ammonia reacts with the acid to form ammonium
sulfate salt while the H2 passes unreacted through the acid [100].
There are several technologies used to separate and purify hydrogen from gas mix-
tures, such as hydrogen-selective membranes or cryogenic distillation for condensing the
impurities and adsorbing them. Alternative technologies for purifying hydrogen at low
energy intensities are pressure-driven membrane processes, an electrochemical membrane
technology based on PEMs with effective hydrogen recovery from H2 /N2 gas mixtures.
Pressure-driven membrane processes are used for hydrogen production because they are
not very energy intensive, and they yield high-purity hydrogen. However, membrane
technologies are advantageous over other purification methods as they commonly depend
on high-pressure feed streams, and hydrogen embrittlement is often experienced. Electro-
chemical hydrogen separation is a one-step operation with the possibility of high separation
efficiency at low cell voltages. It can capture and store CO2 from the feed without the
requirement for any further treatment [149,150].
A commercial process used for a global hydrogen product of 85% is pressure swing
adsorption (PSA), which performs adsorption and desorption at alternating high pressures
and high partial pressures (of 10–40 bar) until reaching the equilibrium pressure loading. It
can remove a large number of impurities corresponding to the difference in adsorption to
desorption loading, at a constant temperature, within few-minute short cycles [151,152].
Due to differences in hydrogen characteristics and conditions, a combination of technologies
(PSA and cryogenic distillation) or a steam-washing scrubber may be required [152,153].
There are several techniques for removing the remaining ammonia from syngas, in-
cluding thermal incineration, which produces NOx, scrubbing, which produces wastewater,
and the catalytic method, which uses several catalysts at temperatures over 650 ◦ C with less
energy efficiency [117,118]. The various organic and inorganic catalysts for the cleanup of
the presence of ammonia in syngas are developed on different materials, such as activated
carbon (2–20 mgNH3 /g sorbent), metal organic frameworks (MOFs-0.27–105 mgNH3 /g
sorbent), covalent organic frameworks (272 mgNH3 /g sorbent), MCM-41 (136 mgNH3 /g
sorbent), alumina (34 mgNH3 /g sorbent), silica gel (85 mgNH3 /g sorbent), and zeolites
(130 mgNH3 /g sorbent) [154]. The zeolite-supported iron (Fe/HZβ) catalyst in the hot-gas
cleanup of remaining ammonia (800 ppm NH3 in the gas mixture) demonstrates better
catalytic activity for ammonia decomposition compared to selective catalytic oxidation
(SCO) techniques at temperatures of 300–550 ◦ C [155]. Inorganic sorbent materials that are
supported with high salt loadings and metal halides are stable such as metal borohydride
and metal fullerides; and can improve ammonia separation from gaseous streams at room
temperature, and improve the partial pressure of ammonia near atmospheric pressure. For
example, carbon silica composites of 2% loaded MgCl2 are capable of separating ammonia
at temperatures above 200 ◦ C [117,118].
The purity of hydrogen is important for PEM fuel cells (PEMFC) and phosphoric
acid fuel cells (PAFC) because of their electrolyte sensitivity to ammonia [149]. Residual
ammonia of more than 0.1 ppm in the fuel stream can damage the catalysts of polymer
electrolyte membrane fuel cells (PEMFC) or decrease the performance of electrochemi-
cal decomposition in aqueous solution due to charge-transfer and diffusion resistances,
resulting in higher tolerance to impurities and better sustainability [53,111,120,156].
Membrane-based separations can simultaneously remove nitrogen at temperatures
higher than 320 ◦ C and provide low-cost separation by coating a single membrane in cheap
material (Pd-coated vanadium). A multifunctional membrane reactor with Pd membrane
walls or a Ru-carbon catalyst at low temperatures can be used for the generation of high-
purity hydrogen with excellent performance [41,66].
Energies 2022, 15, 8246 41 of 49

7. The Challenges and Conclusions


7.1. The Present and Future of Hydrogen and Ammonia Production
Hydrogen production is considered from the view of feedstocks, energy sources,
pathways, and technology in terms of efficiency, cost, and emission. The efficiency of
hydrogen production methods is dependent on their environmental emissions, total cost,
such as the cost of natural gas and electricity, and their energy and exergy efficiencies.
Technologies for hydrogen production require several processes, such as thermochem-
ical, radiochemical, electrochemical, photochemical, biochemical, and integrated systems.
In view of its feedstocks and energy sources, there are pathways for producing various
types of grey, blue and green hydrogen from carbon-based and renewable resources.
Hydrogen is conventionally generated from fossil fuels by reforming, plasma arc decom-
position, coal gasification, and biomass gasification methods, and with high energy and exergy
efficiencies among them, SMR has the highest efficiency, and thus, it is commonly used.
An alternative technology is water electrolysis using electricity power to generate pure
hydrogen with high conversion rates at ambient conditions without the requirement of a
purification system. However, currently, electricity has a high cost and emits GHG.
Pure hydrogen can be produced on a large scale as a by-product of mature industrial
technologies, without additional purification processes, as well as from waste materials,
biochemical processes of microorganisms, and combinations of dark fermentation and
anaerobic digestion. However, these processes have low conversion efficiencies and need
more development.
Ammonia, as a renewable hydrogen carrier, is considered to produce hydrogen due
to its high contents of hydrogen (17.65%) and relative ease of transport and handling
energy. Furthermore, it can supply conversion energy with the utilization of ammonia
combustion. Ammonia is converted into hydrogen under high temperatures in the presence
of a suitable catalyst under standard pressure. Ammonia production can be divided into
three main types (grey, blue, and green) based on its feedstocks and emissions. Grey
ammonia is synthesized through the thermochemical processes, including gasification of
coal, reforming of natural gas, naphtha reforming, pyrolysis, and auto-thermal reactor
(ATR), with high carbon emissions. Blue ammonia with lower carbon is obtained by adding
the carbon capture process. Green ammonia can be produced from renewable feedstocks
such as air and water with sustainable energy sources using water electrolysis instead of
natural gas, oil, or coal.
The Haber–Bosch process is the most common, well-established, economic, and sus-
tainable method of ammonia production. The exergy efficiencies for ammonia production
from natural gas is 65.8% and from biomass-based is 41.3%, with the overall emission range
between 0.5 and 2.3 tCO2 /tNH3 . Biomass gasification has lower efficiency but is cheaper
than natural gas, and it has the additional possibility of using the produced syngas for
electricity generation.
Green ammonia can be produced from renewable feedstock and sustainable energy
sources, such as air and water, by using water electrolysis instead of natural gas, oil, or
coal. However, it is expensive to produce in a green way without carbon dioxide, and its
production process needs more development.

7.2. Challenges of Generating Hydrogen from Ammonia Technologies


Although the utilization of ammonia can play a role in reducing CO2 emissions,
its combustion as fuel produces NOx. Therefore, it is preferred to convert ammonia to
hydrogen, which can be a major carbon-free energy carrier for future energy systems and
can provide sustainable energy for power, industry, buildings, and transport systems.
Table 13 demonstrates some challenges faced in producing hydrogen from ammonia.
The hydrogen production from ammonia requires high temperatures, which can be obtained
by the combustion of carbon-based fuels, and electrical heating, but renewables such as
ammonia or hydrogen are preferred because of costs, availability, and ecofriendly.
Energies 2022, 15, 8246 42 of 49

Table 13. Advantages and disadvantages of using ammonia for hydrogen production.

Advantages Disadvantages
Raw material Renewable ammonia Toxic
Greenhouse gas emissions if using
If using renewable sources
Energy sources carbon-based sources
No greenhouse gas emission
NOx emission if using ammonia as fuel
Clean Large-scale hydrogen production
Output products are green systems are expensive, not
Technology with water emission well-established technology,
Can reduce pollution Energy-intensive, need high
Sustainable temperatures and catalyzer
Efficiency Low overall system efficiency (65%)

The main ammonia-to-hydrogen conversion methods can be summarized based on the


technology, temperature and efficiency, as represented in Table 14. An alternative to thermal
decomposition is the electrochemical method. Furthermore, the radiation energy of an
electron beam, ion beam, plasma, RF, microwave, and solar energy, for the decomposition
of ammonia over the photocatalyst can provide a high hydrogen conversion rate.

Table 14. Comparison of various ammonia decomposition methods for hydrogen production.

Material and Energy


Purification
Requirements Efficiency Temperature
External Heat Catalysts Separation
Thermochemical
X X X High High
process
Electrochemical
× × × High Ambient
process
Photocatalytic × X X Low Ambient

The thermochemical conversion methods of ammonia to hydrogen are considered for


their efficiency, reaction temperature, reactors and catalysts for achieving higher conversion
efficiency under economic conditions.
The selection of a suitable method from the available technology for converting
ammonia to hydrogen requires consideration of the material, temperature and energy
requirements, purification and separation facilities, conversion and energy efficiency, purity,
and yield of hydrogen production.
However, the studies suggested some ways to overcome the challenges throughout
the ammonia decomposition for the hydrogen production process [34,62,80,81]. Therefore,
more efforts and innovations in the technology of ammonia decomposition to hydrogen, are
required to improve energy efficiency, reliability and scalability, which will allow selection
of the most suitable utility systems with minimal energy requirements and lower operating
costs [12,17,43,44,81].

7.3. Efforts for Cost Reduction Technology


Ru-based catalysts are generally used in ammonia cracking for producing hydrogen
as they have the highest catalytic activity in high ammonia concentrations at temperatures
of 425 ◦ C to 500 ◦ C and can reduce emissions ranging from 78% to 95%, but they have the
disadvantages of scarcity and quick deactivation that impede the large-scale application as
an expensive process, resulting in a high cost of ammonia decomposition.
Efforts have been made to find alternative catalysts to replace Ru, which can be
from the various transition metals, including the promoters and supports, such as Fe, Co,
Ni, Ti, V, Mn, and Cr, which have high activity, high stability, weaker nitrogen binding,
effectiveness activity, and low cost.
Energies 2022, 15, 8246 43 of 49

The components of transitional metal include nitrides, carbides, alkali metals, imides
and amides (NaNH2 , Li2 NH). Because of their unique reaction mechanism, fine powder
catalysts of Fe, Ce, Al, Si, Sr, and Zr can be applicable in ammonia decomposition at
relatively low temperatures.
Nano carbon technology can increase catalyst performance, reaction surface area,
reactivity, secured durability, and the size of facilities. The catalytic ability of nanostructured
electro-catalysts usually varies with size and morphology. However, some other limitations
exist in industrial-scale applications.
There are several technologies used to separate and purify hydrogen from remaining
ammonia and nitrogen, such as hydrogen-selective membranes and cryogenic distillation
to condense the impurities and adsorb them.
Alternative technologies for purifying hydrogen at low energy intensities are pressure-
driven membrane processes and electrochemical membrane technologies based on PEMs,
which can generate high-purity hydrogen with high performance and result in a low-cost
separation. However, they are currently used in fuel cells and do not have the perfor-
mance and efficiency required for large-scale hydrogen purification. Although, demand of
hydrogen purity in industrial utilization can be less than that of fuel cells.
However, pressure swing adsorption (PSA) is a commercial process that performs
adsorption and desorption at alternating high pressures and can remove large amounts
of impurities at a constant temperature within a few minutes. Because of differences in
hydrogen characteristics and conditions, a combination of technologies (PSA and cryo-
genic distillation) or the steam-washing scrubber may be required. All of the available
technologies for purifying hydrogen from gas mixtures currently have some advantages
and disadvantages.
Therefore, more efforts and innovations are required in the technologies of ammonia
decomposition to hydrogen to improve their energy efficiency, reliability, and scalability.
These issues can be accomplished by transitioning to hydrogen energy on a large scale,
which would result in a highly efficient energy system with reduced costs.
The review results show that only using hydrogen or ammonia cannot solve the
problem of environmental emissions of fuels, and more efforts are required to improve the
production technology, in addition to using raw low-carbon materials and energy sources
to generate free-carbon hydrogen. Therefore, this paper focuses on ammonia as a renewable
source and provides a suitable cracking method and catalyst for producing green fuel in
the future.

Author Contributions: Conceptualization, D.S.; methodology, D.S.; resources, H.A.Y.R.; data cura-
tion, H.A.Y.R.; writing—original draft preparation, H.A.Y.R.; writing—review and editing, H.A.Y.R.;
visualization, H.A.Y.R.; supervision, D.S.; project administration, D.S.; funding acquisition, D.S. All
authors have read and agreed to the published version of the manuscript.
Funding: This research was funded by KETEP (Korea Institute of Energy Technology Evaluation and
Planning) No. 202003040030090 and KEIT (Korea Evaluation Institute of Industrial Technology) No.
20213030040550.
Data Availability Statement: Not applicable.
Conflicts of Interest: The authors declare no conflict of interest.

References
1. Dincer, I.; Acar, C. Review and evaluation of hydrogen production methods for better sustainability. Int. J. Hydrogen Energy 2015,
40, 11094–11111. [CrossRef]
2. Ye, M.; Sharp, P.; Brandon, N.; Kucernak, A. System-level comparison of ammonia, compressed and liquid hydrogen as fuels for
polymer electrolyte fuel cell powered shipping. Int. J. Hydrogen Energy 2022, 47, 8565–8584. [CrossRef]
3. Zhang, X.; Jiao, K.; Zhang, J.; Guo, Z. A review on low carbon emissions projects of steel industry in the World. J. Clean. Prod.
2021, 306, 127259. [CrossRef]
4. International Energy Agency. Net Zero by 2050: A Roadmap for the Global Energy Sector; International Energy Agency: Paris, France,
2021; 224p, Available online: https://www.iea.org/reports/net-zero-by-2050 (accessed on 20 September 2021).
Energies 2022, 15, 8246 44 of 49

5. Thomas, H. Options for Producing Low-Carbon Hydrogen at Scale; The Royal Society: London, UK, 2018.
6. De Pee, A.; Pinner, D.; Roelofsen, O.; Somers, K. Decarbonization of Industrial Sectors: The Next Frontier; McKensy: Atlanta, GA,
USA, 2018.
7. Hordeski, M.F. Alternative Fuels—The Future of Hydrogen; River Publishers: Gistrup, Denmark, 2020.
8. Eichman, J.; Townsend, A.; Melaina, M. Economic Assessment of Hydrogen Technologies Participating in California Electricity Markets;
National Renewable Energy Lab. (NREL): Golden, CO, USA, 2016.
9. International Energy Agency. The Future of Hydrogen: Seizing Today’s Opportunities; International Energy Agency: Paris, France, 2019.
10. Deloitte, B.E. The Potential of Hydrogen for the Chemical Industry; Deloitte: Beirut, Lebanon, 2021.
11. McKinlay, C.J.; Turnock, S.R.; Hudson, D.A. Route to zero emission shipping: Hydrogen, ammonia or methanol? Int. J. Hydrogen
Energy 2021, 46, 28282–28297. [CrossRef]
12. Andersson, J.; Grönkvist, S. Large-scale storage of hydrogen. Int. J. Hydrogen Energy 2019, 44, 11901–11919. [CrossRef]
13. Hollevoet, L.; De Ras, M.; Roeffaers, M.; Hofkens, J.; Martens, J.A. Energy-Efficient Ammonia Production from Air and Water
Using Electrocatalysts with Limited Faradaic Efficiency. ACS Energy Lett. 2020, 5, 1124–1127. [CrossRef]
14. Forsberg, C. Addressing the low-carbon million-gigawatt-hour energy storage challenge. Electr. J. 2021, 34, 107042. [CrossRef]
15. Park, C.; Koo, M.; Woo, J.; Hong, B.I.; Shin, J. Economic valuation of green hydrogen charging compared to gray hydrogen
charging: The case of South Korea. Int. J. Hydrogen Energy 2022, 47, 14393–14403. [CrossRef]
16. Thomas, J.M.; Edwards, P.P.; Dobson, P.J.; Owen, G.P. Decarbonising energy: The developing international activity in hydrogen
technologies and fuel cells. J. Energy Chem. 2020, 51, 405–415. [CrossRef]
17. Jeerh, G.; Zhang, M.; Tao, S. Recent progress in ammonia fuel cells and their potential applications. J. Mater. Chem. A 2021, 9,
727–752. [CrossRef]
18. Zendrini, M.; Testi, M.; Trini, M.; Daniele, P.; Van Herle, J.; Crema, L. Assessment of ammonia as energy carrier in the use with
reversible solid oxide cells. Int. J. Hydrogen Energy 2021, 46, 30112–30123. [CrossRef]
19. El-Shafie, M.; Kambara, S.; Hayakawa, Y. Hydrogen Production Technologies Overview. J. Power Energy Eng. 2019, 7, 107–154.
[CrossRef]
20. Burandt, T. Analyzing the necessity of hydrogen imports for net-zero emission scenarios in Japan. Appl. Energy 2021, 298, 117265.
[CrossRef]
21. Rogelj, J.; Geden, O.; Cowie, A.; Reisinger, A. Three ways to improve net-zero emissions targets. Nature 2021, 591, 363–365.
[CrossRef] [PubMed]
22. Stangarone, T. South Korean efforts to transition to a hydrogen economy. Clean Technol. Environ. Policy 2020, 23, 509–516.
[CrossRef] [PubMed]
23. Lemmon, E.W.; Bell, I.H.; Huber, M.L.; McLinden, M.O. NIST Standard Reference Database 23: Reference Fluid Thermodynamic and
Transport Properties-REFPROP, Version 10.0; Stand. Ref. Data Program; National Institute of Standards and Technology: Gaithersbg,
MD, USA, 2018.
24. Kobayashi, H.; Hayakawa, A.; Somarathne, K.K.A.; Okafor, E.C. Science and technology of ammonia combustion. Proc. Combust.
Inst. 2019, 37, 109–133. [CrossRef]
25. Sharma, S.; Ghoshal, S.K. Hydrogen the future transportation fuel: From production to applications. Renew. Sustain. Energy Rev.
2015, 43, 1151–1158. [CrossRef]
26. Liu, X.; Elgowainy, A.; Wang, M. Life cycle energy use and greenhouse gas emissions of ammonia production from renewable
resources and industrial by-products. Green Chem. 2020, 22, 5751–5761. [CrossRef]
27. Rivarolo, M.; Riveros-Godoy, G.; Magistri, L.; Massardo, A.F. Clean Hydrogen and Ammonia Synthesis in Paraguay from the
Itaipu 14 GW Hydroelectric Plant. ChemEngineering 2019, 3, 87. [CrossRef]
28. Suleman, F.; Dincer, I.; Agelin-Chaab, M. Environmental impact assessment and comparison of some hydrogen production
options. Int. J. Hydrogen Energy 2015, 40, 6976–6987. [CrossRef]
29. Akarsu, B.; Genç, M.S. Optimization of electricity and hydrogen production with hybrid renewable energy systems. Fuel 2022,
324, 124465. [CrossRef]
30. Mansour-Satloo, A.; Agabalaye-Rahvar, M.; Mirazaei, M.A.; Mohammadi-Ivatloo, B.; Zare, K.; Anvari-Moghaddam, A. A hybrid
robust-stochastic approach for optimal scheduling of interconnected hydrogen-based energy hubs. IET Smart Grid 2021, 4,
241–254. [CrossRef]
31. von Colbe, J.B.; Ares, J.-R.; Barale, J.; Baricco, M.; Buckley, C.; Capurso, G.; Gallandat, N.; Grant, D.M.; Guzik, M.N.; Jacob, I.; et al.
Application of hydrides in hydrogen storage and compression: Achievements, outlook and perspectives. Int. J. Hydrogen Energy
2019, 44, 7780–7808. [CrossRef]
32. Ajanovic, A.; Sayer, M.; Haas, R. The economics and the environmental benignity of different colors of hydrogen. Int. J. Hydrogen
Energy 2022, 47, 24136–24154. [CrossRef]
33. Åhman, M. When Gold Turns to Sand: A Review of the Challenges for Fossil Fuel Rich States Posed by Climate Policy; IMES/EESS Report
No. 124; Lund University: Lund, Sweden, 2021. [CrossRef]
34. Ji, M.; Wang, J. Review and comparison of various hydrogen production methods based on costs and life cycle impact assessment
indicators. Int. J. Hydrogen Energy 2021, 46, 38612–38635. [CrossRef]
35. Kroposki, B.; Levene, J.; Harrison, K.; Sen, P.; Novachek, F. Electrolysis: Information and Opportunities for Electric Power Utilities;
Technical Report NREL/TP-581-40605; NREL: Golden, CO, USA, 2006. [CrossRef]
Energies 2022, 15, 8246 45 of 49

36. Edwards, P.P.; Kuznetsov, V.L.; David, W.I.F. Hydrogen energy. Philos. Trans. R. Soc. A Math. Phys. Eng. Sci. 2007, 365, 1043–1056.
[CrossRef]
37. Uddin, M.N.; Nageshkar, V.V.; Asmatulu, R. Improving water-splitting efficiency of water electrolysis process via highly
conductive nanomaterials at lower voltages. Energy, Ecol. Environ. 2020, 5, 108–117. [CrossRef]
38. Demirbas, A. Future hydrogen economy and policy. Energy Sources, Part B Econ. Planning, Policy 2016, 12, 172–181. [CrossRef]
39. Oni, A.; Anaya, K.; Giwa, T.; Di Lullo, G.; Kumar, A. Comparative assessment of blue hydrogen from steam methane reforming,
autothermal reforming, and natural gas decomposition technologies for natural gas-producing regions. Energy Convers. Manag.
2022, 254, 115245. [CrossRef]
40. Fakeeha, A.; Ibrahim, A.A.; Aljuraywi, H.; Alqahtani, Y.; Alkhodair, A.; Alswaidan, S.; Abasaeed, A.E.; Kasim, S.O.; Mahmud, S.;
Al-Fatesh, A.S. Hydrogen Production by Partial Oxidation Reforming of Methane over Ni Catalysts Supported on High and Low
Surface Area Alumina and Zirconia. Processes 2020, 8, 499. [CrossRef]
41. Kumar, S.S.; Himabindu, V. Hydrogen production by PEM water electrolysis—A review. Mater. Sci. Energy Technol. 2019, 2,
442–454. [CrossRef]
42. Ghavam, S.; Vahdati, M.; Wilson, I.A.G.; Styring, P. Sustainable Ammonia Production Processes. Front. Energy Res. 2021, 9, 580808.
[CrossRef]
43. Jung, S.-C.; Chung, K.-H.; Choi, J.; Park, Y.-K.; Kim, S.-J.; Kim, B.-J.; Lee, H. Photocatalytic hydrogen production using liquid
phase plasma from ammonia water over metal ion-doped TiO2 photocatalysts. Catal. Today 2022, 397–399, 165–172. [CrossRef]
44. Obata, K.; Kishishita, K.; Okemoto, A.; Taniya, K.; Ichihashi, Y.; Nishiyama, S. Photocatalytic decomposition of NH3 over TiO2
catalysts doped with Fe. Appl. Catal. B Environ. 2014, 160–161, 200–203. [CrossRef]
45. Acar, C.; Dincer, I. Review and evaluation of hydrogen production options for better environment. J. Clean. Prod. 2019, 218,
835–849. [CrossRef]
46. Wen, D.; Aziz, M. Flexible operation strategy of an integrated renewable multi-generation system for electricity, hydrogen,
ammonia, and heating. Energy Convers. Manag. 2022, 253, 115166. [CrossRef]
47. Manna, J.; Jha, P.; Sarkhel, R.; Banerjee, C.; Tripathi, A.; Nouni, M. Opportunities for green hydrogen production in petroleum
refining and ammonia synthesis industries in India. Int. J. Hydrogen Energy 2021, 46, 38212–38231. [CrossRef]
48. Cha, J.; Park, Y.; Brigljević, B.; Lee, B.; Lim, D.; Lee, T.; Jeong, H.; Kim, Y.; Sohn, H.; Mikulčić, H.; et al. An efficient process for
sustainable and scalable hydrogen production from green ammonia. Renew. Sustain. Energy Rev. 2021, 152, 111562. [CrossRef]
49. Kanoglu, M.; Dincer, I.; Rosen, M.A. Understanding energy and exergy efficiencies for improved energy management in power
plants. Energy Policy 2007, 35, 3967–3978. [CrossRef]
50. El-Shafie, M.; Kambara, S.; Hayakawa, Y. Energy and exergy analysis of hydrogen production from ammonia decomposition
systems using non-thermal plasma. Int. J. Hydrogen Energy 2021, 46, 29361–29375. [CrossRef]
51. Acar, C.; Dincer, I. 3.1 Hydrogen Production. Compr. Energy Syst. 2018, 3, 1–40.
52. Colakoglu, M.; Durmayaz, A. Energy, exergy and economic analyses and multiobjective optimization of a novel solar multi-
generation system for production of green hydrogen and other utilities. Int. J. Hydrogen Energy 2022, 47, 19446–19462. [CrossRef]
53. Lim, D.-K.; Plymill, A.B.; Paik, H.; Qian, X.; Zecevic, S.; Chisholm, C.R.; Haile, S.M. Solid Acid Electrochemical Cell for the
Production of Hydrogen from Ammonia. Joule 2020, 4, 2338–2347. [CrossRef]
54. Esily, R.R.; Chi, Y.; Ibrahiem, D.M.; Chen, Y. Hydrogen strategy in decarbonization era: Egypt as a case study. Int. J. Hydrogen
Energy 2022, 47, 18629–18647. [CrossRef]
55. International Energy Agency. CCUS in Clean Energy Transitions; International Energy Agency: Paris, France, 2020.
56. Wu, X.-Y.; Luo, Y.; Hess, F.; Lipiński, W. Editorial: Sustainable Hydrogen for Energy, Fuel and Commodity Applications. Front.
Energy Res. 2021, 9, 698669. [CrossRef]
57. Al-Breiki, M.; Bicer, Y. Comparative life cycle assessment of sustainable energy carriers including production, storage, overseas
transport and utilization. J. Clean. Prod. 2020, 279, 123481. [CrossRef]
58. Boerner, L.K. Industrial ammonia production emits more CO2 than any other chemical-making reaction. Chemists want to change
that. Chem. Eng. News 2019, 97, 1–9.
59. International Fertilizer Industry Association. Fertilizers, Climate Change and Enhancing Agricultural Productivity Sustainably;
International Fertilizer Industry Association: Paris, France, 2009.
60. Worrell, E.; Boyd, G. Bottom-up estimates of deep decarbonization of U.S. manufacturing in 2050. J. Clean. Prod. 2021, 330, 129758.
[CrossRef]
61. Keçebaş, A.; Kayfeci, M. Chapter 1—Hydrogen Properties. In Solar Hydrogen Production; Calise, F., D’Accadia, M.D., Santarelli, M.,
Lanzini, A., Ferrero, D., Eds.; Academic Press: Cambridge, MA, USA, 2019; pp. 3–29.
62. Hasan, M.H.; Mahlia, T.M.I.; Mofijur, M.; Fattah, I.M.R.; Handayani, F.; Ong, H.C.; Silitonga, A.S. A Comprehensive Review on
the Recent Development of Ammonia as a Renewable Energy Carrier. Energies 2021, 14, 3732. [CrossRef]
63. Chu, K.H.; Lim, J.; Mang, J.S.; Hwang, M.-H. Evaluation of strategic directions for supply and demand of green hydrogen in
South Korea. Int. J. Hydrogen Energy 2021, 47, 1409–1424. [CrossRef]
64. Yuksel, Y.E.; Ozturk, M.; Dincer, I. Design and analysis of a new solar hydrogen plant for power, methane, ammonia and urea
generation. Int. J. Hydrogen Energy 2022, 47, 19422–19445. [CrossRef]
Energies 2022, 15, 8246 46 of 49

65. Lee, K.; Liu, X.; Vyawahare, P.; Sun, P.; Elgowainy, A.; Wang, M. Techno-economic performances and life cycle greenhouse gas
emissions of various ammonia production pathways including conventional, carbon-capturing, nuclear-powered, and renewable
production. Green Chem. 2022, 24, 4830–4844. [CrossRef]
66. Aziz, M.; Wijayanta, A.T.; Nandiyanto, A.B.D. Ammonia as Effective Hydrogen Storage: A Review on Production, Storage and
Utilization. Energies 2020, 13, 3062. [CrossRef]
67. Bird, F.; Clarke, A.; Davies, P.; Surkovic, E. Ammonia: Fuel and Energy Store; KBR Inc.: Houston, TX, USA, 2020.
68. Valera-Medina, A.; Gutesa, M.; Xiao, H.; Pugh, D.; Giles, A.; Goktepe, B.; Marsh, R.; Bowen, P. Premixed ammonia/hydrogen
swirl combustion under rich fuel conditions for gas turbines operation. Int. J. Hydrog. Energy 2019, 44, 8615–8626. [CrossRef]
69. Flórez-Orrego, D.; Maréchal, F.; Junior, S.D.O. Comparative exergy and economic assessment of fossil and biomass-based routes
for ammonia production. Energy Convers. Manag. 2019, 194, 22–36. [CrossRef]
70. IEA. Ammonia Technology Roadmap; International Energy Agency: Paris, France, 2021.
71. MacFarlane, D.R.; Cherepanov, P.V.; Choi, J.; Suryanto, B.H.; Hodgetts, R.Y.; Bakker, J.M.; Vallana, F.M.F.; Simonov, A.N. A
Roadmap to the Ammonia Economy. Joule 2020, 4, 1186–1205. [CrossRef]
72. IEA; ICCA. Technology Roadmap—Energy and GHG Reductions in the Chemical Industry via Catalytic Processes; International Energy
Agency: Paris, France, 2013.
73. Ikäheimo, J.; Kiviluoma, J.; Weiss, R.; Holttinen, H. Power-to-ammonia in future North European 100% renewable power and
heat system. Int. J. Hydrogen Energy 2018, 43, 17295–17308. [CrossRef]
74. Nguyen, T.; Abdin, Z.; Holm, T.; Mérida, W. Grid-connected hydrogen production via large-scale water electrolysis. Energy
Convers. Manag. 2019, 200, 112108. [CrossRef]
75. Ouikhalfan, M.; Lakbita, O.; Delhali, A.; Assen, A.H.; Belmabkhout, Y. Toward Net-Zero Emission Fertilizers Industry: Greenhouse
Gas Emission Analyses and Decarbonization Solutions. Energy Fuels 2022, 36, 4198–4223. [CrossRef]
76. Ausfelder, F.; Herrmann, E.O.; González, L.F.L. Perspective Europe 2030 Technology Options for CO2 —Emission Reduction of Hydrogen
Feedstock in Ammonia Production; Dechema: Frankfurt, Germany, 2022.
77. Penkuhn, M.; Tsatsaronis, G. Comparison of different ammonia synthesis loop configurations with the aid of advanced exergy
analysis. Energy 2017, 137, 854–864. [CrossRef]
78. Ghannadzadeh, A.; Sadeqzadeh, M. Diagnosis of an alternative ammonia process technology to reduce exergy losses. Energy
Convers. Manag. 2016, 109, 63–70. [CrossRef]
79. Flórez-Orrego, D.; de Oliveira Junior, S. Modeling and optimization of an industrial ammonia synthesis unit: An exergy approach.
Energy 2017, 137, 234–250. [CrossRef]
80. Wu, S.; Salmon, N.; Li, M.M.-J.; Bañares-Alcántara, R.; Tsang, S.C.E. Energy Decarbonization via Green H2 or NH3 ? ACS Energy
Lett. 2022, 7, 1021–1033. [CrossRef]
81. Mukherjee, S.; Devaguptapu, S.V.; Sviripa, A.; Lund, C.R.; Wu, G. Low-temperature ammonia decomposition catalysts for
hydrogen generation. Appl. Catal. B Environ. 2018, 226, 162–181. [CrossRef]
82. Chai, W.S.; Bao, Y.; Jin, P.; Tang, G.; Zhou, L. A review on ammonia, ammonia-hydrogen and ammonia-methane fuels. Renew.
Sustain. Energy Rev. 2021, 147, 111254. [CrossRef]
83. Fowler, D.; Coyle, M.; Skiba, U.; Sutton, M.A.; Cape, J.N.; Reis, S.; Sheppard, L.J.; Jenkins, A.; Grizzetti, B.; Galloway, J.N.; et al.
The global nitrogen cycle in the twenty-first century. Philos. Trans. R. Soc. B Biol. Sci. 2013, 368, 20130164. [CrossRef]
84. Lin, Q.; Jiang, Y.; Liu, C.; Chen, L.; Zhang, W.; Ding, J.; Li, J. Instantaneous hydrogen production from ammonia by non-thermal
arc plasma combining with catalyst. Energy Rep. 2021, 7, 4064–4070. [CrossRef]
85. Chiuta, S.; Everson, R.C.; Neomagus, H.W.; van der Gryp, P.; Bessarabov, D.G. Reactor technology options for distributed
hydrogen generation via ammonia decomposition: A review. Int. J. Hydrogen Energy 2013, 38, 14968–14991. [CrossRef]
86. Bell, T.E.; Murciano, L.T. H2 Production via Ammonia Decomposition Using Non-Noble Metal Catalysts: A Review. Top. Catal.
2016, 59, 1438–1457. [CrossRef]
87. Ross, J.R. The Kinetics and Mechanisms of Catalytic Reactions. In Contemporary Catalysis; Elsevier: Amsterdam, The Netherlands,
2019; pp. 161–186. [CrossRef]
88. Slycke, J.; Mittemeijer, E.; Somers, M. Thermodynamics and kinetics of gas and gas–solid reactions. In Thermochemical Surface
Engineering of Steels; Woodhead Publishing: Sawston, UK, 2015; pp. 3–111. [CrossRef]
89. Wang, W.; Padban, N.; Ye, Z.; Andersson, A.A.; Bjerle, I. Kinetics of Ammonia Decomposition in Hot Gas Cleaning. Ind. Eng.
Chem. Res. 1999, 38, 4175–4182. [CrossRef]
90. Kiełbasa, K.; Pelka, R.; Arabczyk, W. Studies of the Kinetics of Ammonia Decomposition on Promoted Nanocrystalline Iron Using
Gas Phases of Different Nitriding Degree. J. Phys. Chem. A 2010, 114, 4531–4534. [CrossRef] [PubMed]
91. Kashkarov, S.; Li, Z.; Molkov, V. Blast wave from a hydrogen tank rupture in a fire in the open: Hazard distance nomograms.
Int. J. Hydrog. Energy 2012, 45, 2429–2446. [CrossRef]
92. Hu, Z.-P.; Weng, C.-C.; Chen, C.; Yuan, Z.-Y. Catalytic decomposition of ammonia to COx-free hydrogen over Ni/ZSM-5 catalysts:
A comparative study of the preparation methods. Appl. Catal. A Gen. 2018, 562, 49–57. [CrossRef]
93. Ojelade, O.A.; Zaman, S.F. Ammonia decomposition for hydrogen production: A thermodynamic study. Chem. Pap. 2020, 75,
57–65. [CrossRef]
94. Chein, R.-Y.; Chen, Y.-C.; Chang, C.-S.; Chung, J. Numerical modeling of hydrogen production from ammonia decomposition for
fuel cell applications. Int. J. Hydrogen Energy 2010, 35, 589–597. [CrossRef]
Energies 2022, 15, 8246 47 of 49

95. Baek, S.-H.; Yun, K.; Kang, D.-C.; An, H.; Park, M.; Shin, C.-H.; Min, H.-K. Characteristics of High Surface Area Molybdenum
Nitride and Its Activity for the Catalytic Decomposition of Ammonia. Catalysts 2021, 11, 192. [CrossRef]
96. Itoh, N.; Oshima, A.; Suga, E.; Sato, T. Kinetic enhancement of ammonia decomposition as a chemical hydrogen carrier in
palladium membrane reactor. Catal. Today 2014, 236, 70–76. [CrossRef]
97. Li, G.; Kanezashi, M.; Yoshioka, T.; Tsuru, T. Ammonia decomposition in catalytic membrane reactors: Simulation and experimen-
tal studies. AIChE J. 2012, 59, 168–179. [CrossRef]
98. Xie, T.; Xia, S.; Jin, Q. Thermodynamic Optimization of Ammonia Decomposition Solar Heat Absorption System Based on
Membrane Reactor. Membranes 2022, 12, 627. [CrossRef]
99. Do, S.-H.; Roh, J.S.; Park, H.B. Carbon-free Hydrogen Production Using Membrane Reactors. Membr. J. 2018, 28, 297–306.
[CrossRef]
100. Barisano, D.; Canneto, G.; Nanna, F.; Villone, A.; Fanelli, E.; Freda, C.; Grieco, M.; Lotierzo, A.; Cornacchia, G.; Braccio, G.; et al.
Investigation of an Intensified Thermo-Chemical Experimental Set-Up for Hydrogen Production from Biomass: Gasification
Process Integrated to a Portable Purification System—Part II. Energies 2022, 15, 4580. [CrossRef]
101. Abashar, M.; Al-Sughair, Y.; Al-Mutaz, I. Investigation of low temperature decomposition of ammonia using spatially patterned
catalytic membrane reactors. Appl. Catal. A Gen. 2002, 236, 35–53. [CrossRef]
102. Li, G.; Yu, X.; Yin, F.; Lei, Z.; Zhang, H.; He, X. Production of hydrogen by ammonia decomposition over supported Co3O4
catalysts. Catal. Today 2022, 402, 45–51. [CrossRef]
103. Chen, Y.-L.; Juang, C.-F.; Chen, Y.-C. The Effects of Promoter Cs Loading on the Hydrogen Production from Ammonia Decompo-
sition Using Ru/C Catalyst in a Fixed-Bed Reactor. Catalysts 2021, 11, 321. [CrossRef]
104. Cechetto, V.; Di Felice, L.; Medrano, J.A.; Makhloufi, C.; Zuniga, J.; Gallucci, F. H2 production via ammonia decomposition in a
catalytic membrane reactor. Fuel Process. Technol. 2021, 216, 106772. [CrossRef]
105. Abashar, M. Ultra-clean hydrogen production by ammonia decomposition. J. King Saud Univ. Eng. Sci. 2018, 30, 2–11. [CrossRef]
106. Zhang, Z.; Liguori, S.; Fuerst, T.F.; Way, J.D.; Wolden, C.A. Efficient Ammonia Decomposition in a Catalytic Membrane Reactor To
Enable Hydrogen Storage and Utilization. ACS Sustain. Chem. Eng. 2019, 7, 5975–5985. [CrossRef]
107. Cechetto, V.; Di Felice, L.; Martinez, R.G.; Plazaola, A.A.; Gallucci, F. Ultra-pure hydrogen production via ammonia decomposition
in a catalytic membrane reactor. Int. J. Hydrogen Energy 2022, 47, 21220–21230. [CrossRef]
108. Heidenreich, S.; Müller, M.; Foscolo, P.U. Chapter 5—Advanced Process Combination Concepts. In Advanced Biomass Gasification;
Heidenreich, S., Müller, M., Foscolo, P.U., Eds.; Academic Press: Cambridge, MA, USA, 2016; pp. 55–97.
109. Yang, X.; Wang, S.; He, Y. Review of catalytic reforming for hydrogen production in a membrane-assisted fluidized bed reactor.
Renew. Sustain. Energy Rev. 2021, 154, 111832. [CrossRef]
110. Mazzone, S.; Campbell, A.; Zhang, G.; García-García, F. Ammonia cracking hollow fibre converter for on-board hydrogen
production. Int. J. Hydrogen Energy 2021, 46, 37697–37704. [CrossRef]
111. Wan, Z.; Tao, Y.; Shao, J.; Zhang, Y.; You, H. Ammonia as an effective hydrogen carrier and a clean fuel for solid oxide fuel cells.
Energy Convers. Manag. 2020, 228, 113729. [CrossRef]
112. Makepeace, J.W.; He, T.; Weidenthaler, C.; Jensen, T.R.; Chang, F.; Vegge, T.; Ngene, P.; Kojima, Y.; de Jongh, P.E.; Chen, P.; et al.
Reversible ammonia-based and liquid organic hydrogen carriers for high-density hydrogen storage: Recent progress. Int. J.
Hydrogen Energy 2019, 44, 7746–7767. [CrossRef]
113. Lucentini, I.; Garcia, X.; Vendrell, X.; Llorca, J. Review of the Decomposition of Ammonia to Generate Hydrogen. Ind. Eng. Chem.
Res. 2021, 60, 18560–18611. [CrossRef]
114. Seyfeli, R.C.; Varisli, D. Ammonia decomposition reaction to produce COx-free hydrogen using carbon supported cobalt catalysts
in microwave heated reactor system. Int. J. Hydrogen Energy 2020, 45, 34867–34878. [CrossRef]
115. Chen, C.; Wu, K.; Ren, H.; Zhou, C.; Luo, Y.; Lin, L.; Au, C.; Jiang, L. Ru-Based Catalysts for Ammonia Decomposition: A
Mini-Review. Energy Fuels 2021, 35, 11693–11706. [CrossRef]
116. Yuzawa, H.; Mori, T.; Itoh, H.; Yoshida, H. Reaction Mechanism of Ammonia Decomposition to Nitrogen and Hydrogen over
Metal Loaded Titanium Oxide Photocatalyst. J. Phys. Chem. C 2012, 116, 4126–4136. [CrossRef]
117. Akkerman, Q.A.; Manna, L. What Defines a Halide Perovskite? ACS Energy Lett. 2020, 5, 604–610. [CrossRef]
118. Assirey, E.A.R. Perovskite synthesis, properties and their related biochemical and industrial application. Saudi Pharm. J. 2019, 27,
817–829. [CrossRef]
119. Hill, A.K.; Torrente-Murciano, L.T. Low temperature H2 production from ammonia using ruthenium-based catalysts: Synergetic
effect of promoter and support. Appl. Catal. B Environ. 2015, 172–173, 129–135. [CrossRef]
120. Lamb, K.E.; Dolan, M.D.; Kennedy, D.F. Ammonia for hydrogen storage; A review of catalytic ammonia decomposition and
hydrogen separation and purification. Int. J. Hydrogen Energy 2019, 44, 3580–3593. [CrossRef]
121. Lente, G. Comment on “‘Turning Over’ Definitions in Catalytic Cycles”. ACS Catal. 2013, 3, 381–382. [CrossRef]
122. Turnover Frequency—An Overview|ScienceDirect Topics. Available online: https://www.sciencedirect.com/topics/chemistry/
turnover-frequency (accessed on 13 September 2022).
123. Chehade, G.; Dincer, I. Progress in green ammonia production as potential carbon-free fuel. Fuel 2021, 299, 120845. [CrossRef]
124. Jiang, K.; Li, K.; Liu, Y.-Q.; Lin, S.; Wang, Z.; Wang, D.; Ye, Y. Nickel-cobalt nitride nanoneedle supported on nickel foam as an
efficient electrocatalyst for hydrogen generation from ammonia electrolysis. Electrochimica Acta 2021, 403, 139700. [CrossRef]
Energies 2022, 15, 8246 48 of 49

125. Mazzone, S.; Goklany, T.; Zhang, G.; Tan, J.; Papaioannou, E.I.; García-García, F. Ruthenium-based catalysts supported on carbon
xerogels for hydrogen production via ammonia decomposition. Appl. Catal. A Gen. 2022, 632, 118484. [CrossRef]
126. Yin, S.; Xu, B.; Zhu, W.; Ng, C.; Zhou, X.; Au, C. Carbon nanotubes-supported Ru catalyst for the generation of COx-free hydrogen
from ammonia. Catal. Today 2004, 93–95, 27–38. [CrossRef]
127. Yin, S.-F.; Xu, B.-Q.; Ng, C.-F.; Au, C.-T. Nano Ru/CNTs: A highly active and stable catalyst for the generation of COx-free
hydrogen in ammonia decomposition. Appl. Catal. B Environ. 2004, 48, 237–241. [CrossRef]
128. El-Shafie, M.; Kambara, S.; Hayakawa, Y. Development of zeolite-based catalyst for enhancement hydrogen production from
ammonia decomposition. Catal. Today 2021, 397–399, 103–112. [CrossRef]
129. Liu, Y.; Wang, H.; Yuan, X.; Wu, Y.; Wang, H.; Tan, Y.Z.; Chew, J.W. Roles of sulfur-edge sites, metal-edge sites, terrace sites, and
defects in metal sulfides for photocatalysis. Chem Catal. 2021, 1, 44–68. [CrossRef]
130. Podila, S.; Driss, H.; Ali, A.M.; Al-Zahrani, A.A.; Daous, M.A. Influence of Ce substitution in LaMO3 (M = Co/Ni) perovskites for
COx-free hydrogen production from ammonia decomposition. Arab. J. Chem. 2021, 15, 103547. [CrossRef]
131. Pinzón, M.; Sánchez-Sánchez, A.; Sánchez, P.; de la Osa, A.; Romero, A. Ammonia as a carrier for hydrogen production by using
lanthanum based perovskites. Energy Convers. Manag. 2021, 246, 114681. [CrossRef]
132. Qiu, Y.; Fu, E.; Gong, F.; Xiao, R. Catalyst support effect on ammonia decomposition over Ni/MgAl2O4 towards hydrogen
production. Int. J. Hydrogen Energy 2021, 47, 5044–5052. [CrossRef]
133. Pinzón, M.; Romero, A.; de Lucas-Consuegra, A.; de la Osa, A.R.; Sánchez, P. COx -free hydrogen production from ammonia at
low temperature using Co/SiC catalyst: Effect of promoter. Catal. Today 2022, 390, 34–47.
134. Gholami, Z.; Tišler, Z.; Rubáš, V. Recent advances in Fischer-Tropsch synthesis using cobalt-based catalysts: A review on supports,
promoters, and reactors. Catal. Rev. 2021, 63, 512–595. [CrossRef]
135. Jackson, C.; Fothergill, K.; Gray, P.; Haroon, F.; Makhloufi, C.; Kezibri, N.; Davey, A.; Hote, O.L.; Zarea, M.; Davenne, T.; et al.
Ammonia to Green Hydrogen Project; Feasibility Study; Ecuity: Birmingham, UK, 2020; pp. 1–70.
136. Wang, Z.; Yao, Y.; Chen, R.; Wang, Z. Research progress on electrocatalytic decomposition of ammonia for hydrogen production.
CIESC J. 2022, 73, 1008–1021. Available online: https://hgxb.cip.com.cn (accessed on 15 March 2022).
137. McEnaney, J.M.; Singh, A.R.; Schwalbe, J.A.; Kibsgaard, J.; Lin, J.C.; Cargnello, M.; Jaramillo, T.F.; Nørskov, J.K. Ammonia
synthesis from N2 and H2 O using a lithium cycling electrification strategy at atmospheric pressure. Energy Environ. Sci. 2017, 10,
1621–1630. [CrossRef]
138. Gao, W.; Guo, J.; Wang, P.; Wang, Q.; Chang, F.; Pei, Q.; Zhang, W.; Liu, L.; Chen, P. Production of ammonia via a chemical looping
process based on metal imides as nitrogen carriers. Nat. Energy 2018, 3, 1067–1075. [CrossRef]
139. Hargreaves, J.S.J. Nitrides as ammonia synthesis catalysts and as potential nitrogen transfer reagents. Appl. Petrochem. Res. 2014,
4, 3–10. [CrossRef]
140. Giddey, S.; Badwal, S.; Kulkarni, A. Review of electrochemical ammonia production technologies and materials. Int. J. Hydrogen
Energy 2013, 38, 14576–14594. [CrossRef]
141. Kyriakou, V.; Garagounis, I.; Vasileiou, E.; Vourros, A.; Stoukides, M. Progress in the Electrochemical Synthesis of Ammonia.
Catal. Today 2017, 286, 2–13. [CrossRef]
142. Seyfeli, R.C.; Varisli, D. Performance of microwave reactor system in decomposition of ammonia using nickel based catalysts
with different supports. Int. J. Hydrogen Energy 2022, 47, 15175–15188. [CrossRef]
143. Zhang, S.; He, Z.; Zuoli, H.; Zhang, J.; Zang, Q.; Wang, S. Building heterogeneous nanostructures for photocatalytic ammonia
decomposition. Nanoscale Adv. 2020, 2, 3610–3623. [CrossRef] [PubMed]
144. Chang, F.; Gao, W.; Guo, J.; Chen, P. Emerging Materials and Methods toward Ammonia-Based Energy Storage and Conversion.
Adv. Mater. 2021, 33, 2005721. [CrossRef] [PubMed]
145. Elysabeth, T.; Mulia, K.; Ibadurrohman, M.; Dewi, E.L.; Slamet. A comparative study of CuO deposition methods on titania
nanotube arrays for photoelectrocatalytic ammonia degradation and hydrogen production. Int. J. Hydrogen Energy 2021, 46,
26873–26885. [CrossRef]
146. Hayakawa, Y.; Miura, T.; Shizuya, K.; Wakazono, S.; Tokunaga, K.; Kambara, S. Hydrogen production system combined with a
catalytic reactor and a plasma membrane reactor from ammonia. Int. J. Hydrogen Energy 2019, 44, 9987–9993. [CrossRef]
147. Giddey, S.; Badwal, S.P.S.; Munnings, C.; Dolan, M. Ammonia as a Renewable Energy Transportation Media. ACS Sustain. Chem.
Eng. 2017, 5, 10231–10239. [CrossRef]
148. Kojima, Y. Hydrogen storage materials for hydrogen and energy carriers. Int. J. Hydrogen Energy 2019, 44, 18179–18192. [CrossRef]
149. Vermaak, L.; Neomagus, H.W.J.P.; Bessarabov, D.G. Hydrogen Separation and Purification from Various Gas Mixtures by Means
of Electrochemical Membrane Technology in the Temperature Range 100–160 ◦ C. Membranes 2021, 11, 282. [CrossRef]
150. Huang, F.; Pingitore, A.T.; Benicewicz, B.C. Electrochemical Hydrogen Separation from Reformate Using High-Temperature
Polybenzimidazole (PBI) Membranes: The Role of Chemistry. ACS Sustain. Chem. Eng. 2020, 8, 6234–6242. [CrossRef]
151. Du, Z.; Liu, C.; Zhai, J.; Guo, X.; Xiong, Y.; Su, W.; He, G. A Review of Hydrogen Purification Technologies for Fuel Cell Vehicles.
Catalysts 2021, 11, 393. [CrossRef]
152. Speight, J.G. Chapter 15—Hydrogen Production. In Heavy Oil Recovery and Upgrading; Gulf Professional Publishing: Houston, TX,
USA, 2019; pp. 657–697.
153. Yáñez, M.; Relvas, F.M.; Ortiz, A.; Gorri, D.; Mendes, A.; Ortiz, I. PSA purification of waste hydrogen from ammonia plants to
fuel cell grade. Sep. Purif. Technol. 2019, 240, 116334. [CrossRef]
Energies 2022, 15, 8246 49 of 49

154. Malmali, M.; Le, G.; Hendrickson, J.; Prince, J.; McCormick, A.V.; Cussler, E.L. Better Absorbents for Ammonia Separation. ACS
Sustain. Chem. Eng. 2018, 6, 6536–6546. [CrossRef]
155. Çetin, Y.; Sarioğlan, A.; Okutan, H. Comparison of catalytic activities both for selective oxidation and decomposition of ammonia
over FE/HZB catalyst. J. Turk. Chem. Soc. Sect. A Chem. 2016, 4, 227. [CrossRef]
156. Wassie, S.A.; Medrano, J.A.; Zaabout, A.; Cloete, S.; Melendez, J.; Tanaka, D.A.P.; Amini, S.; Annaland, M.V.S.; Gallucci, F.
Hydrogen production with integrated CO2 capture in a membrane assisted gas switching reforming reactor: Proof-of-Concept.
Int. J. Hydrogen Energy 2018, 43, 6177–6190. [CrossRef]

You might also like